You are on page 1of 8

International Journal of Machine Tools & Manufacture 46 (2006) 651658

www.elsevier.com/locate/ijmactool

Determination of temperature distribution in the cutting zone


using hybrid analytical-FEM technique
W. Grzesik*
Department of Manufacturing Engineering and Production Automation, Faculty of Mechanical Engineering, Technical University of Opole, P.O. Box 321,
45-271 Opole, Poland
Received 4 April 2005; accepted 5 July 2005
Available online 9 September 2005

Abstract
In this study, the temperature distribution in the cutting zone was determined by integrating thermal analytical and simulation models of
orthogonal cutting process with uncoated and coated carbide tools. Primarily, 2D FEM simulations were run to provide numerical solutions
of temperatures occurring at different points through the chip/tool contact region and the coating/substrate boundary under defined cutting
conditions. In addition, an analytical model for heat transfer in the cutting tool and its partitioning, proposed in References [W. Grzesik, P.
Nieslony, Physics based modelling of interface temperatures in machining with multilayer coated tools at moderate cutting speeds, Int. J.
Mach. Tools Manufact. 44 (2004) 889901; W. Grzesik, P. Nieslony, A computational approach to evaluate temperature and heat partition in
machining with multilayer coated tools, Int. J. Mach. Tools Manufact. 43 (2003) 13111317], was employed to generate the input data to
computations of the toolchip interface temperature. The changes of the temperature distribution fields resulting from varying heat flux
transfer conditions are the main findings of the FEM simulations. Finally, the analytically and numerically predicted average temperatures
were validated against the tool-work thermocouple-based measurements and discussed in terms of relevant literature data.
q 2005 Elsevier Ltd. All rights reserved.
Keywords: Machining; Steels; Temperature distribution; Hybrid modelling

1. Introduction
It is a special kind of technical paradox that relatively
little effort has been directed toward measuring temperature
distribution in the tool, which is correlated directly with its
wear rate. As a substitute various simulation models which
are capable of predicting the tool temperature distribution
with satisfactory accuracy are developed. This trend is
motivated by the strong belief that the substantial
improvement of metal removal efficiency can also be
achieved by more sophisticated modelling of these
processes at a system level, that means by generation of
the house of models [1].
For the past 50 years, metal cutting researchers have
developed many modelling techniques including analytical
techniques, slip-line solutions, empirical approaches
* Tel.: C48 77 4006290; fax: C48 77 4006342.
E-mail address: grzesik@polo.po.opole.pl.

0890-6955/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmachtools.2005.07.009

and finite element techniques. In recent years, the finite


element method has particularly become the main tool for
simulating metal cutting processes [24]. Finite element
models are widely used for calculating the stress, strain,
strain-rate and temperature distributions in the primary,
secondary and tertiary sub-cutting zones. In consequence,
temperatures in the tool, chip and workpiece, as well as
cutting forces, plastic deformation (shear angle and chip
thickness), chip formation and possibly its breaking can be
determined faster than by using costly and time consuming
experiments.
Typical approaches for numerical modelling of metal
cutting processes are Lagrangian and Eulerian techniques, as
well as a combination of both called an arbitrary Lagrangian
Eulerian formulation (denoted in the literature by ALE
acronym) [2,5]. It should be noticed that all these methods
are mathematically equivalent. The major feature of
Lagrangian formulation used in this study is that the
discretized mesh is attached to the workpiece and the material
model is elasticplastic, only plastic, or viscoplastic.
Early finite element analyses were performed by Usui
and Shirakashi [6], Iwata et al. [7] and Strenkowski

652

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

and Carroll [8], who were the first to use Eulerian


formulations for steady-state metal cutting simulations. On
the other hand, Marusich and Ortiz [9] have developed a
Lagrangian formulation in which the material model
contains deformation hardening, thermal softening and
strain-rate sensitivity tightly coupled with a transient heat
conduction analysis appropriate for finite deformations.
Most of the early investigations on the FEM modelling of
orthogonal cutting were limited by the assumption of a
perfectly sharp cutting edge which satisfies the fundamental
2D cutting model proposed by Ernst and Merchant.
In recent years, a few trials were undertaken [10] to
extend the FEM modelling technique to real non-sharp
cutting tool geometries including round/hone and T-land/
chamfer edges. In particular, the Lagrangian thermoviscoplastic cutting simulation of 0.2% carbon steel
performed by Yen et al. [10] have revealed that plastic
deformation and interface temperature change considerably
by increasing the edge radius of the hone tools.
It is specially important that FEM analysis can help to
investigate some thermodynamical effects occurring in the
cutting zone which, so far, cannot be measured directly [3].
An example for such effects is the influence of cutting tool
coatings on the heat transfer and friction, and resulting
cutting temperature distribution in the chip and the tool.
Moreover, it can explain the controversial thermal barrier
(thermal insulation) effect provided by multilayer coatings
with an intermediate Al2O3 ceramic layer with extremely
low thermal conductivity at the contact temperature higher
than 1000 K [1113].
It should be noted that the majority of previous
orthogonal metal cutting simulations is devoted to uncoated
carbide tools and now the opposite trend to consider single
and multiply coatings has been observed [6,1315]. In fact,
the first completed work focusing on the evaluation of a
predictive orthogonal cutting model for coated carbide tools
with multiple coating layers using the FEM was presented
by Yen et al. [13]. In this model, the thermal properties of
the three layer-TiC/Al2O3/TiN were implemented by
considering individual coating layers and a composite
monolayer coating with substitute (equivalent) thermal
conductivity and diffusivity. On the other hand, the
comparison with the experimental data from the literature
[16] has also been performed. Separate Lagrangian and
Eulerian calculations were used to obtain the steady-state
solution of temperature in the tool and chip/workpiece. The
main finding from this simulation was that for AISI 1045
steel the predicted steady-state interface temperatures are in
good agreement with the experimental values given in Ref.
[16] within 511% difference. Moreover, for cutting
velocity vcZ220 m/min and feed rate fZ0.16 mm/rev, the
composite layer approach which replaces TiC/Al2O3/TiN
coating has predicted slightly higher temperatures of about
30 8C.
Marusich et al. have documented in Ref. [15] that a
significant reduction in temperature which occurs for

the multiply coated tool causes that the carbide substrate


is on the order of 100 8C cooler than the uncoated reference.
Unfortunately, the coefficient of friction was not changed
numerically when coating was added. In this case,
simulations were performed with Third Wave Systems
AdvantEdge finite element based modelling software [17].
In this paper, a Lagrangian finite element code
AdvantEdge was also applied to construct a coupled
thermo-mechanical finite element model of plane-strain
orthogonal metal cutting with continuous chip formation
produced by plane-faced uncoated and differently coated
carbide tools.
The entire cutting process is simulated, i.e. from the
initial to the steady-state phase. The workpiece material of
choice, AISI 1045 carbon steel, is modelled as thermo
elasticplastic, while the flow stress is considered to be a
function of strain, strain-rate and temperature to represent
better the real behaviour in cutting. The constant coefficient
of friction between the tool and chip is based on the
Coulomb friction.

2. Finite element model


2.1. Graphical representation of the simulation model
As mentioned above the finite element model used for the
plane-strain orthogonal metal cutting simulation is based on
the Lagrangian techniques and an explicit dynamic, thermomechanically coupled modelling software with adaptive
remeshing. This means that the initial mesh becomes
distorted after a certain length of cut as shown in Fig. 1b
and is remeshed in this vicinity to form a regular mesh
again. The upper part of mesh, which constitutes the
removed workpiece material, is finer, to enable the stress,
strain, strain-rate and temperature in the chip and the tool tip
regime to be accurately predicted. In the original model
proposed [15] constant thickness coating layers are added to
the tool substrate to create multiple coating as shown in
Fig. 1a. For the dimensions of the rest, a coarser mesh is
sufficient. In the case shown in Fig. 1b, the workpiece
consists of about 1340 six-nodded plane-strain triangular
elements and 1460 nodes. Dimensions of the element size
can range from the minimum value of 0.02 mm to maximum
one of 0.1 mm.
On the other hand, the tool model consists of the
adequate number of node planar heat-transfer elements,
because heat transfer analysis is carried out for it. The lower
half of its mesh, expected to be in contact with the chip, is
modelled with a finer mesh, in order to be able to predict the
temperature field developed in the tool.
2.2. Workpiece material modelling
The workpiece material used for the plain-strain
orthogonal metal cutting simulation is C45 steel of

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

653

Fig. 1. The mesh model for a TiC/Al2O3/TiN coated tool used in [13] (a) and shape of the deformed chip after a tool path of 4.0 mm with cutting edge radius of
33 mm and 1460 nodes obtained in the present study (b).

the ultimate tensile strength UTSZ670 MPa. The material


is modelled as isotropic elasticplastic with isotropic strainhardening. In the cutting processes, the deformation of the
material in the primary and secondary cutting zones occurs
at elevated temperatures and very high strains and strain
rates (105107 sK1). On the other hand, the remainder of the
workpiece deforms at moderate or even low strain rates.
Consequently, Thirdwave AdvantEdge computes the
increase in flow stress due to the strain rate sensitivity
using the following formula [15,18]:


3_p 1=m1
p
s Z sf 3 1 C p
(1)
3_o
where s is the effective von Misses stress, sf is the flow
stress, 3p is the accumulated plastic strain, 3po is a reference
plastic strain rate, m1 is the strain rate sensitivity exponent.
The flow stress is determined using a power hardening
law model with thermal softening effect, namely


3p 1=n
sf Z so $QT$ 1 C p
(2)
3o
where so is the initial yield stress at the reference
temperature 3po, is the reference plastic strain, n is the
hardening exponent and Q(T) is the thermal softening factor
ranging from 1 at ambient to O at melt.
2.3. Heat transfer
As commonly accepted the two main sources of heat in
cutting processes are the plastic work and the dissipation of
friction energy at the toolchip interface, which are
practically fully converted into heat. Usually, the percentage
of the cutting power converted into heat is assumed to be
equal to 90% [19,20]. The portion of frictional work being
converted into heat is taken as 1.0 in this study. Because
coatings influence, the heat partition at the interface the
proper values of the heat partition coefficient, which
strongly depend on the contact temperature, are incorporated in this simulation. On the other hand, coatings can

significantly change chip formation and heat transfer at the


toolchip interface and moderate tool wear [11,16]. In this
study, the coating has linear elastic mechanical but nonlinear, temperature-dependent thermal properties
2.4. Boundary conditions and process parameters
The tool geometry and the cutting conditions used for the
orthogonal metal cutting simulation are presented in
Table 1.
Machining is performed at ambient temperature assuming the initial temperature of both workpiece and the tool is
equal to 20 8C. According the simulation model proposed
previously by Yen et al. [13] the heat losses to the
environment from the non-contact surfaces of the tool and
workpiece due to convection are determined by the
distributed heat flux as
q Z hTw KTo

(3)

where the convection heat transfer coefficient of the


workpiece material hZ20 W/m2 8C [13], TW is the workpiece temperature and T0 is room temperature.

3. Thermal analytical model


In general, the modelling concept proposed is based on
the well-known principle of the simultaneous action of two
independent heat sources, which suggests that the total heat
Table 1
Cutting conditions and tool geometry
Tool geometry
Tool rake angle
Tool clearance angle
Measured cutting edge radius
Cutting conditions
Undeformed chip thickness
Width of cut
Cutting speed
Coulomb friction coefficient

K58
58
33 mm
0.16 mm
2.0 mm
103.2 m/min
0.5

654

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

flux is generated by aggregation of the plastic deformation


and sliding friction effects. There are the shear zone
(primary deformation zone-PDZ) and frictional heat sources
(secondary deformation zone-SDZ) [21].
3.1. Calculation of equivalent thermal properties
In the composite layer concept, in which all components
are represented by one proportionally thicker monolayer,
the equivalent (effective, substitute) thermal conductivity
depends on the thickness of each component and the number
of coating layers. It can be determined using the expression
well-known in thermodynamics [10], as follows
t
P

xi

leq

x1 x2
x
C C/C t
l1 l2
lt

(4)

where xi is the thickness of the selected i-layer (in this case


iZ3), li is thermal conductivity of i-layer, Sxi is total
thickness of the stack (composite layer) and leq the
equivalent thermal conductivity of the total layer, t denotes
the top layer.
The equivalent thermal diffusivity can be determined as
the ratio of the equivalent thermal conductivity to the
equivalent volumetric heat capacity (Ceq). The proper
formula can be derived by summing volumes of the
individual layers Vi to obtain the total coating, as follows [22]
t
P
xi Ci
ceq Z 1 t
(5a)
P
xi
1

aeq Z

leq
Ceq

(5b)

where Ci, cpi and ri are the volumetric heat capacity, the
specific heat and density of i-layer, respectively.
Calculated values of equivalent conductivity and
diffusivity for the cutting speed of 103.2 m/min
(at corresponding measured cutting temperature of
592.3 8C) are equal to leqZ17.51 W/m K and aeqZ
5.89!10-6 m2/s.
3.2. Determination of the heat partition coefficient
and temperature components
In this modelling case study, two expressions for the heat
partition coefficient proposed by Shaw-RSH [11] and
Reznikov-RR [14] are used. The expression for RSH
coefficient is
1
p
(6)
RSH Z
1 C 0:754lTeq =lW =Aa NT 
In Eq. (6), thermal conductivities of work (lW) and
tool (lTZlTeq) materials, thermal number (NT) and area

shape factor (A) describe the thermal and geometrical


features of the interface. Adequate formulae for NT and A
are provided in [14].
The expression for coefficient RR implements the
velocity and duration of frictional heat source represented
here by Peclet and Fourier numbers, respectively. Therefore, the heat partitioning can be determined from the
following relation [14]
RR Z

1
p
1 C 3lTeq =2lW aW =aTeq 

(7)

where lTeq is the equivalent thermal conductivity of the


composite layer and a Teq is corresponding thermal
diffusivity.
The average interface temperature is defined as the sum
of the mean shear-plane temperature (Q s ) and the mean
temperature rise due to friction (DQ f ), namely [22]:
Q t Z Q s C DQ f

(8)

The computation algorithm which considers the mean


temperature rise due to friction and plastic deformation is
provided in the original form in Ref. [22].

4. Results and discussion


In this study, numerical simulations of the heat transfer in
the cutting zone by conduction, including the tool and the
chip, were carried out for constant cutting parameters, i.e.
cutting speed of 103.2 m/min, feed rate of 0.16 mm/rev and
depth of cut of 2 mm and the tools geometrical features,
listed in Table 1.
4.1. Results of numerical simulations
4.1.1. Intensity of heat sources
As shown in Figs. 2a and b, the heat rate increases in the
vicinity of the cutting edge and changes for the cutting tool
materials used. The maximum values of volumetric cutting
energy equal to 17,742 W/mm3 was computed for a P20
uncoated carbide (Fig. 2a), whereas for a three layer coating
it reaches 19,447 W/mm3 as presented in Fig. 2b.
It should be noted that FEM programme used automatically scales the heat rate and it was not possible to establish
accurately the area of heat penetration for the cases
considered in this investigation.
4.1.2. Isotherm patterns in the cutting zone
The temperature distribution in the workpiece, tool and
chip closer to the cutting edge after a tool path of 4.0 mm,
is shown in the form of the magnified images in Figs. 3b
and 4b. The maximum temperatures located in these
selected zones determined from the thermal maps
presented in Figs. 3a and 4a and those computed
analytically are listed in Table 2.

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

655

Fig. 2. Distribution of heat sources for ISO P20 uncoated (a) and TiC/Al2O3/TiN (b) coated carbide tools. Cutting conditions as in Table 1.

It should be noted that the maximum temperatures existing


in the tool determined by FEM and analytical method differ
substantially for both cutting tool materials of about 150 8C.
All simulations performed have revealed that cutting tool
coatings tested influenced distinctly the performance and
intensity of the thermal interactions when turning C45 carbon
steel. In particular, coatings change both the heat transfer and
its distribution in the cutting zone as shown in Fig. 4b. In
comparison to an uncoated P20 carbide tool (Fig. 3b) the
coating applied caused that the fraction of heat which flows
into the tool decreases.
On the other hand, it can be observed that more heat is
transferred to the chip and workpiece. Moreover, coatings
cause that areas with the maximum temperatures are
localized near the chip and workpiece. In consequence,
the maximum interface temperature exists in the vicinity of
the cutting edge, i.e. in the first part of the toolchip contact.
This finding agrees qualitatively with FDM predictions
documented in [23]. This effect is especially visible for

the three-layer TiC/Al2O3/TiN coating as shown in Fig. 4b.


Furthermore, it was revealed based on temperature
distribution illustrated in Fig. 4b that temperature developing on the workpiece surface increases about 50 8C in
comparison to the conventional tool used.
It is evident that this effect can be related to differences in
the thermal properties of the tool materials. In particular, the
thermal conductivity of Al2O3 ceramic layer in the
TiC/Al2O3/TiN coating decreases distinctly and it is
apparent that at higher contact temperatures the carbide
substrate is partly thermally insulated by the coating. As a
result, heat flow into the substrate is more difficult than for
the coatings including TiC, TiN and TiCN layers [16].
4.2. Results of analytical predictions
Fig. 5 shows the results of analytical predictions of the
average interface temperature for uncoated and multilayer
coated tools using the heat partitioning estimated by means

Fig. 3. Isotherm patterns in ISO P20 uncoated carbide tools (a) and magnification of temperature distribution in the vicinity of the cutting edge (b). Cutting
conditions as in Table 1.

656

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

Fig. 4. Isotherm patterns in TiC/Al2O3/TiN coated carbide tools (a) and magnification of temperature distribution in the vicinity of the cutting edge (b). Cutting
conditions as in Table 1.

of Eqs. (6) and (7). Moreover, the estimated measurement


errors resulting from both calibration and computerized data
processing are comprehensively listed in Table 3. It should
be pointed out that the cutting temperatures were
determined with corresponding maximum individual percentage deviations of G33.4 and G49 8C for uncoated and
coated tools, respectively.
Fig. 5 conforms the applicability of the analytical
thermal models proposed and shows good results in light
of the thermocouple measurements. In such a confrontation
the absolute errors determined for the reference cutting
speed of 103.2 are in the range of 10-200C and tend to
decrease further for higher cutting speeds.
4.3. FEM prediction of cutting temperature

estimation of friction when using the Coulomb friction


model.
4.4. Validation of the FEM and analytical predictions
The comparison between measured and computed values
of the average toolchip interface temperature is presented
in Fig. 7.
Primarily, it was observed that the FEM code used
overestimated analytical predictions and consequently
employed inadequate values of the friction coefficient. For
this reason, the simulation results obtained were validated
by the comparison to appropriate toolwork thermocouple
measurements provided in [16]. For a 10 mm equivalent
800

A-P20
M-P20
700

Temperature,C

666.7C
648.9C

A-3L
M-3L
592.3C

600

581.1C

500
103.2 m/min

The appropriate temperature histories obtained during


simulations along with corresponding values of the average
toolchip interface temperatures are shown in Fig. 6a and b.
It was established based on temperature traces with tool
travel that the time required to reach the steady-state
temperature was 0.350.60 ms depending on the tool
material used. For example, the value of the average contact
temperature of about 680 8C was computed for P20
uncoated tools and as expected a lower temperature of
650 8C was fixed for the equivalent composite layer of
10 mm thickness. It is suggested that these differences in the
contact temperatures can be referred to inadequate
Table 2
Values of maximum temperature in the tool and chip (in 8C)
Coating/substrate material
P20
TiC/Al2O3/
TiN-P20

Tool
Analytical
885.3
777.7

Chip
FEM
635.3
524.5

FEM
679.2
650.6

400
40

60

80

100

120

140

160

180

200

220

Cutting speed, m/min


Fig. 5. The average temperature vs. cutting speed. Symbols: A, analytical
prediction; M, measurement; 3L, substitute coating.

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

657

Table 3
Specification of values of percentage deviations for Figs. 5 and 7
Cutting speed (m/min)
Insert type
Average temperature (8C)
Estimated error range (8C)
Insert type
Average temperature (8C)
Estimated error range (8C)

51.37
62.34
72.24
P20 uncoated
587.62
608.49
624.92
G28.70
G29.30
G25.40
coated withTiC/Al2O3/TiN
547.00
559.15
568.62
G41.00
G43.30
G44.10

103.2

124.69

144.48

178.13

206.4

649.09
G30.60

666.70
G33.00

690.07
G30.50

708.88
G33.40

736.56
G28.90

756.72
G27.80

582.39
G45.80

592.31
G48.00

605.34
G49.00

615.71
G48.50

630.82
G44.00

641.69
G43.20

600

600

500

500

400
300

400
300

200

200

100

100
0

0
0.0000

650.6C

(b) 700

679.2C

Temperature,C

Temperature,C

(a) 700

89.06

0.0005

0.0010

0.0015

0.0020

0.0025

0.0005

0.001

0.0015

0.002

0.0025

Time, s

Time, s

Fig. 6. Average interface temperature trace vs. simulation time for uncoated (a) and three-layer coated (b) carbide tools. Cutting conditions as in Table 1.

5. Conclusions
In this paper, a coupled thermo-mechanical model of
orthogonal metal cutting was used to determine the
distribution of temperature in the tool and the chip as
well as the average toolchip interface temperature.
Moreover, this modelling issue was extended to the
analytical prediction of the average toolchip temperature incorporating equivalent thermal properties for the
multilayer coating applied and corresponding heat
partitioning.
It is shown that the outcomes of the FEM and analytical
models provide quite satisfactory and physically
supported results, for both uncoated and three-layer
coated tools, concerning values and distributions of
cutting temperatures. However, a better accuracy can
probably be obtained by tuning friction parameter and
heat partition to coated tools with real thicknesses (in
this study they were based on the manufacturers data
not on measurements using for instance scanning
microscopy).
It is apparent that the key assumption of the constant
friction coefficient of 0.5 (Coulomb friction) in a FEM
model is not appropriate for machining with coated tools

for which friction can be more intensive. According to


Refs. [20] and [24] the proper values of m determined
experimentally from forces data are equal to 0.63 and
0.68 for C45-P20 and C45-TiC/Al2O3/TiN-P20 couples,
respectively.
The analytical thermal model using an equivalent onelayer coating predicts the average steady-state temperatures at the toolchip interface within the percentage
accuracy of about 2.7 and 1.9% for uncoated ISO P20
carbide and three-layer coated tools, respectively. In
contrast, these accuracies obtained in the finite element
simulation are of 1.9 and 10%. This fact evidently
confirms the need for supplying realistic coating
800
measurement
700

Average interface
temperature, C

coating, this divergence was found to be about 10%, and for


uncoated tools both predictions provided comparable
accuracy of 35%.

simulation FEM

666.7 679.2

648.9

analytical model

592.3 650.6
581.1

600
500
400
300
200
100
0

P20
L3
Cutting tool material

Fig. 7. Comparison between measured and computed values of the average


tool-chip interface temperature. Cutting conditions as in Table 1. Symbols:
L3-TiC/Al2O3/TiN coating, P20-ISO conventional carbide grade.

658

W. Grzesik / International Journal of Machine Tools & Manufacture 46 (2006) 651658

properties and heat partitioning data for FEM simulation


cutting model.
The finite element simulations performed yield the
evidence of existence and localization of the secondary
shear zone. In particular, it was documented that for
coated tools areas with the maximum temperatures are
localized near the chip and workpiece. Also the substrate
under the thin coating is visibly cooler in comparison to
uncoated tools.

References
[1] C.A. van Luttervelt, T.H.C. Childs, I.S. Jawahir, F. Klocke,
P.K. Venuvinod, Present situation and future trends in modelling of
machining operations, Ann. CIRP 47/2 (1998) 587626.
[2] S.M. Athavale, J.S. Strenkowski, Finite element modelling of
machining: from proof-of-concept to engineering applications,
Mach. Sci. Technol. 2 (2) (1998) 317342.
[3] F. Klocke, T. Beck, S. Hoppe, T. Krieg, et al., Examples of FEM
application in manufacturing technology, J. Mater. Proc. Technol. 120
(2002) 450457.
[4] W. Zhang, O.W. Dillon, I.S. Jawahir, A finite element analysis of 2D machining with a grooved tool, Trans. NAMRI/SME 29 (2001)
327334.
[5] M.R. Movahhedy, M.S. Gadala, Y. Altintas, Simulation of chip
formation in orthogonal metal cutting process: an ALE finite element
approach, Mach. Sci. Technol. 4 (1) (2000) 1542.
[6] E. Usui, T. Shirakashi, Mechanics of Machining-from Descriptive to
Predictive Theory On the Art of Cutting Metals-75 Years Later, vol. 7,
ASME PED, New York, 1982. pp.1355.
[7] K. Iwata, K. Osakada, Y. Teresaka, Process modeling of orthogonal
cutting by a rigid-plastic finite element method, J. Eng. Mat. Technol.
106 (1984) 132138.
[8] J.S. Strenkowski, J.T. Carroll III, A. finite, element model of
orthogonal metal cutting, J. Eng. Ind. 107 (1985) 347354.
[9] T.D. Marusich, M. Ortiz, Modelling and simulation of high-speed
machining, Int. J. Numer. Methods Eng. 38 (1995) 36753694.

[10] Y.-Ch Yen, A. Jain, T. Altan, A finite element analysis of orthogonal


machining using different tool edge geometries, J. Mater. Proc.Technol. 146 (2004) 7281.
[11] W. Grzesik, The role of coatings in controlling the cutting process
when turning with coated indexable inserts, J. Mater. Proc. Technol.
37 (1998) 133143.
[12] J. Rech, A. Kusiak, J.L. Battaglia, Tribological and thermal functions
of cutting tool coatings, Surf. Coat. Technol. 186 (2004) 364371.
[13] Y.-Ch. Yen, A. Jain, P. Chigurupati, W.-T. Wu, T. Altan, Computer
Simulation of Orthogonal Cutting Using a Tool With Multiple
Coatings Proceedings of the 6th CIRP International Workshop on
Modeling of Machining Operation, McMaster University, ON,
Canada, May 20, 2003 pp. 119130.
[14] J. Leopold, M. Meisel, R. Wohlgemuth, J. Liebich, High
performance computing of coating-substrate systems, Surf. Coat.
Technol. 142144 (2001) 916922.
[15] T.D. Marusich, C.J. Brand, J.D. Thiele, A Methodology for
Simulation of Chip Breakage in Turning Processes Using an
Orthogonal Finite Element Model Proceedings of the 5th CIRP
International Workshop on Modeling of Machining Operation, West
Lafayette, IN, USA, May 2021, 2002 pp. 139148.
[16] W. Grzesik, An integrated approach to evaluating the tribo-contact for
coated cutting inserts, Wear 240 (2000) 918.
[17] Third Wave Systems AdvantEdgee Users manual, Version 4.2,
2003.
[18] H. Bil, S.E. Kilic, A.E. Tekkaya, A comparison of orthogonal cutting
data from experiments with three different finite element models, Int.
J. Mach. Tools Manuf. 44 (2004) 933944.
[19] A.G. Mamalis, M. Horvath, A.S. Branis, D.E. Manolakos, Finite
element simulation of chip formation in orthogonal metal cutting,
J. Mater. Proc. Technol. 110 (2001) 1927.
[20] Ch. Shet, X. Deng, Finite element analysis of the orthogonal metal
cutting process, J. Mater. Proc. Technol. 105 (2000) 95109.
[21] M.C. Shaw, Metal Cutting Principles, Clarendon Press, Oxford, 1989.
[22] W. Grzesik, P. Nieslony, Physics based modelling of interface
temperatures in machining with multilayer coated tools at moderate
cutting speeds, Int. J. Mach. Tools Manufact. 44 (2004) 889901.
[23] W. Grzesik, M. Bartoszuk, P. Nieslony, Finite difference analysis of
the thermal behaviour of coated tools in orthogonal cutting of steels,
Int. J. Mach. Tools Manufact. 44 (2004) 14511462.
[24] W. Grzesik, P. Nieslony, Prediction of friction and heat flow in
machining incorporating thermophysical properties of the coatingchip interface, Wear 256 (2004) 108117.

You might also like