You are on page 1of 15

International Journal of Mechanical Sciences 43 (2001) 2699–2713

Finite element simulation of chip ow in metal machining


M.H. Dirikolua; ∗ , T.H.C. Childsb , K. Maekawac
a
Department of Mechanical Engineering, University of Kirikkale, Yahsihan 71450, Kirikkale, Turkey
b
Department of Mechanical Engineering, University of Leeds, Leeds LS2 9JT, UK
c
Department of Mechanical Engineering, Ibaraki University, Hitachi, Japan

Received 2 May 2000; received in revised form 6 February 2001

Abstract
Finite element studies of machining are becoming ever more sophisticated. A basic approach which
removes the need, in an elastic–plastic analysis, to follow the development of chip formation from initial
contact between work and tool, is the iterative convergence method (ICM). It develops a steady-state
chip formation from an initial state of a fully formed chip loaded against a tool. It relies for its accuracy
on the assumption that its simpli3ed loading path coincides with the real developed ow at the end of
the simulation. This paper examines the robustness of this assumption by studying the sensitivity of the
simulation to changes of detail, within the ICM method, of how the ow develops; and it compares the
simulated results with experiments. The experiment involves the turning of three free cutting steels, for
which experimental ow stress variations with strain, strain rate and temperature, as well as information
about the friction interaction between chip and tool, are available. The changes to the simulation method
considered here are the structure of the 3nite element mesh, the measures of judging when the ow
is fully developed, how the chip separates from the work at the cutting edge and the friction laws
used during the approach to fully developed ow. It is shown that these do a6ect the outcomes of the
simulation but within the ranges studied only to a minor extent and good agreement with experiment is
achieved. ? 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Metal machining; Finite elements; Iterative convergence method; Free-cutting steels

1. Introduction

Simulation is an important tool for capturing the changing nature of manufacturing industry.
The competitive world economy requires that, for an industry to survive, it must minimise
response times and costs, as well as maximise the e:ciency and quality in producing a product.


Corresponding author. Fax: +90-318-357-459.
E-mail address: dirikolu@kku.edu.tr (M.H. Dirikolu).

0020-7403/01/$ - see front matter ? 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 0 - 7 4 0 3 ( 0 1 ) 0 0 0 4 7 - 9
2700 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

Nomenclature
B; M; N; k ∗ ; m∗ parameters in Eqs. (1) and (2)
Cp speci3c heat
K thermal conductivity
U chip velocity
V cutting speed
k shear ow stress
m friction factor
n friction law exponent (Eq. (3d))
t1 undeformed chip thickness
t2 chip thickness
 rake angle
C equivalent strain
Ċ equivalent strain rate
 shear plane angle
 friction coe:cient
C equivalent stress
n rake face normal contact stress (positive compressive)

Metal machining is a widely used and costly manufacturing process, to which these considera-
tions apply. It has attracted great attention for a long time: one strong focus is what mechanisms
a6ect chip formation, as this is where most of the energy is expended during this process.
Better control of chip handling, tool wear, dimensional accuracy, and surface integrity are other
underlying concerns. Trial and error methods for investigations of all these are time consuming
and costly. A more e6ective way is to model and then simulate the process. Many machining
models have been proposed [1– 4]. More recently, advances in computer processing power, have
led to numerically based models, using the 3nite element method [5 –8].
This paper is mainly concerned with advances in a 3nite element simulation of chip formation,
originating from studies by Usui and his group [9,10], and known as the iterative convergence
method (ICM). It was initially implemented on a supercomputer but was later simpli3ed and
pre- and post-processors added for interactive use on a CAD computer [11–13], and is cur-
rently running on personal computers. Separate but collaborative developments in Japan and
the UK (at Leeds) have led to di6erences in principle in certain areas of the model (here-
after the two versions will be called simulation A and simulation B, respectively). The paper
reports the di6erences between the versions: automatic mesh generation and model discretisation,
the treatment of chip separation at the cutting edge, the program termination criteria, friction
boundary conditions, and material thermal property modelling. These as well as the theory of
the simulation model are discussed in Section 2. The simulations require two sets of input data
(in addition to tool and work thermal properties). These are the workpiece material ow stress
behaviour and the cutting tool–chip interface friction stress data: these have been, respectively,
obtained from split-Hopkinson bar [14] and split-tool dynamometer [15,16] tests. Cutting tests
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2701

have also been carried out in order to establish a common base for comparing both simulation
programs.
Material property information concerning the work and tool materials, needed for the sim-
ulations, is presented in Section 3. The results of the cutting tests (cutting and thrust forces;
and mean contact temperatures between the chip and the tool) and a comparison with simulated
results are in Section 4. The two simulation variants do yield di6erent results: these are discussed
in Section 5; and an overall conclusion is made in Section 6.

2. Simulation of machining

2.1. Basic considerations

Machining processes involve large plastic deformations accompanied by elastic, thermal, and
frictional e6ects in a region enclosed by the workpiece, the chip, and the cutting tool. Due to the
coupling of these e6ects, any attempt to model this process by the 3nite element method should
consider the geometric and material non-linearities due to large deformations and the mutual
dependence of the ow stress and friction characteristics through the 3elds of the temperature,
strain, and strain rate. Dealing with these aspects by modelling, at least in elastic–plastic cases,
necessitates an incremental straining approach to be taken. An incremental approach can simulate
the whole machining process from beginning to end [6].
The ICM approach supposes that if the steady state of continuous chip formation is what is of
interest, then the transition stage at the beginning of cutting may be neglected. It suggests that
the simulation may be started by loading a tool against an already formed chip and stopped
when the chip ow reaches its steady state. This requires very little tool displacement, and
vastly saves on computing time and cost: for example, computing time on a PC with a Pentium
II processor takes 5 –10 min, as opposed to maybe 5 h if chip formation is followed from the
3rst contact between tool and work. The issues are two-fold: the shape of the already formed
chip is not known a priori, and hence an iteration to 3nd that shape is needed; and it is an
assumption that the simpli3ed work and chip loading path will come into coincidence with the
actual loading path. It is the second issue that has driven the work reported here.

2.2. Theory and methodology of the simulations

A ow chart showing the steps involved in the ICM is given in Fig. 1. At the start of the
calculation, a chip is supposed already to exist on the work. It is assumed to be straight, with
a simple at primary shear plane, as illustrated in Fig. 2. A guess is made of the shear plane
angle and a 3nite element mesh is generated for this shape: how this is done is described in
more detail in Section 2.3. The work material is given its initial properties (also see Section
2.3). The work is then displaced towards the tool, assumed to be rigid, with displacement
and friction boundary conditions again described more in Section 2.3. An elastic–plastic 3nite
element analysis follows the development of the stress 3eld and the velocity distribution until
the cutting forces and plastic ow zone reach their steady state. In the elastic–plastic analysis,
the material is assumed to obey the Prandtl–Reuss ow rules and the von Mises yield criterion.
2702 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

Fig. 1. Flow chart for the ICM methodology.

Fig. 2. Outline of the initially assumed chip geometry.


M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2703

The use of the latter is not universally accepted for metal cutting modelling in the sense that
this criterion assumes independence of compressive normal stress, while there have been many
arguments regarding the e6ect of this on the shear stress on the shear plane, such as those quoted
in Refs. [3,17–19]. The element sti6ness equations are formulated for large strain increments in
an updated Lagrangian approach, including element shape change and rotation terms, and the
relaxation of over-constraint of incompressibility [20,21].
Typically, the calculated ow does not match the initially assumed ow lines. The obtained
velocity distribution at the 3nal stage of the elastic–plastic calculation is used to reform the ow
lines in the deformed workpiece, to reduce the di6erence between the actual and the assumed
ow. By integrating the strain rates along the modi3ed ow lines, the equivalent strain of
each element is calculated. Next, the thermal 3nite element module calculates the temperature
distributions in the work and the tool. This is achieved by solving the thermal di6usion equation
in its variational form, which has been used previously in machining analyses [22]. The new
estimates of temperatures, strains, and strain rates are used to modify the current ow stress.
This is done iteratively until the temperature 3eld converges. If the stress, strain, velocity and
temperature 3elds thus obtained are not satisfactorily close to those obtained in the previous
ICM iteration, the elastic–plastic 3nite element calculation is repeated after pulling back the tool
to its original position and resetting the components of stress and strain to zero: the elements in
the work and chip are given the updated ow stress, work-hardening characteristics, temperatures
and nodal coordinates. The iteration is continued until a converged state is reached.

2.3. Di5erences—and common ground—between the simulations

2.3.1. The 7nite element mesh


In both simulations A and B, the mesh consists of three-node constant strain triangular
elements, organised along the ow streamlines. This type of element is known to behave more
rigidly than a real continuum in the fully plastic range [20]: both simulations deal with this
problem by considering the triangles in pairs, e6ectively creating a four-node element that does
not su6er from this problem [21]. However, in the case of simulation A, the elements are
re3ned, in an e:cient manner, just in the regions of primary and secondary shear (Fig. 3), in
a hand-generated manner. Simulation B uses an automatic mesh generator (a result is shown
in Fig. 4), by which it is easier to vary the re3nement with, for example, change of cutting
speed that in uences the temperature gradients in the ow 3eld; but it is less e:cient in that
re3nement is also carried out in some areas where it is not needed.

2.3.2. Chip=work separation criteria


A particular feature of simulating machining is to model the chip separation at the cutting
edge. Commonly, in Lagrangian analyses, a small crack is introduced at the cutting edge (Fig.
5a). In models that truly follow the chip loading path, the issue arises, that is, when to propagate
the crack from one element to the next, as the tool moves through the work: critical displace-
ment, strain and strain energy criteria have been used. In the ICM method, the tool displacement
required to reach the fully loaded state is typically only 5% of the uncut chip thickness. If the
crack tip node does not reach the cutting edge within the simulation’s displacement, there is
no need for a crack propagation criterion: the size of the element at the cutting edge is chosen
2704 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

Fig. 3. (a) The 3nite element meshing scheme—simulation A; (b) magni3ed mesh detail around the cutting
edge—simulation A.

Fig. 4. (a) The 3nite element meshing scheme—simulation B; (b) magni3ed mesh detail around the cutting
edge—simulation B.
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2705

Fig. 5. Chip separation treatments: (a) a crack (simulation A); and (b) a singular element (simulation B) at the
cutting edge.

to ensure this. This is the approach taken in simulation A. In practice, in the continuum real
world, there may be no crack at the cutting edge. In simulation B, a singular element has
been introduced into the work at the cutting edge (Fig. 5b). The element is initially triangular
with two coincident nodes at the cutting edge. As the work moves relative to the tool, one of
the nodes moves up the rake face and the other along the free cut surface. This approach has
been used successfully before, for ow separation at the corner of a backward-extrusion punch,
by Mori et al. [23]. Modi3cation of the sti6ness matrix for the singular element follows the
treatment by Becker et al. [24]

2.3.3. Convergence criteria


The use of an iterative solution method requires appropriate convergence (or termination)
criteria. If an inappropriate criterion is used, the iteration may be terminated before the necessary
solution accuracy is reached or be continued after the required accuracy has already been reached
[25]. The ICM method requires three convergence=termination criteria: one for the elastic–plastic
analysis (loop I in Fig. 1), a second for the thermal analysis (loop II), and a third for ICM
termination (loop III).
Within loop I, with increasing displacement of the work towards the tool, the load on the
tool 3rst rises and elements originally elastic become plastic in the primary and secondary shear
zones but, after some displacement, it is observed that the load starts to fall and plastically loaded
elements return to an elastic state: this is because the actual loading path is not the same as
in steady-state chip formation. Simulation A terminates the tool displacement when the load
reaches its maximum value or when elements start to unload, whichever happens 3rst. Another
way of terminating the displacement, employed in simulation B, is to consider the development
of nodal velocities with increasing displacement between the work and tool. The ow of work
towards the tool is t1 V . The ow of the chip away from the primary shear zone is t2 U , where
U is the chip velocity, as shown in Fig. 2. It is observed that as displacement increases, so
does the ratio of t2 U to t1 V . Eventually, it either reaches a value of unity (indicating volume
2706 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

conservation of the ow) or passes through a maximum at a value less than unity: the tool
displacement is terminated at either of these occurrences (in the 3nally converged state—loop
III—the ratio is always close to unity).
In loop II, both simulations use four iterations to terminate the thermal calculations. Temper-
ature convergence always occurs within this number. The convergence of loop III is checked
on the basis of similarity of chip shapes, forces, stress and temperature distributions between
two consecutive iterations.

2.3.4. Material 9ow properties


It is well accepted that the in uence of strain, strain rate and temperature on work ow
stress must be included in successful simulations of machining: many empirical equations have
been developed to describe this [26]. Both simulations in this work use the empirical equation
structure developed by Usui et al. [14], based on split Hopkinson bar hot compression testing.
The full expression for ow stress, including strain path e6ects, is
 M  −m∗   −m∗ =N N
Ċ ∗ 
Ċ  ∗ 
Ċ 
C = B ek T e−k T=N d C ; (1)
1000 1000  strainpath 1000 

where the coe:cients B; M and N re ect the ow stress at a strain rate of 1000=s and a strain
of 1, the strain-rate sensitivity, and strain hardening index, respectively, and are functions of
temperature T . k ∗ and m∗ are constants associated with strain path dependence. In the absence
of strain path e6ects (k ∗ = m∗ = 0), Eq. (1) reduces to the following form:
 M
Ċ
C = B C N : (2)
1000
Simulation A has used the full Eq. (1), while simulation B has used Eq. (2).

2.3.5. Material thermal properties


In simulation A, thermal conductivity of the work and tool materials and speci3c heat of the
work have been assumed independent of temperature, while in simulation B the work material
conductivity and speci3c heat has been allowed to vary linearly with temperature.

2.3.6. Friction boundary conditions


Realistic characterisation of the friction interaction between the chip and the tool is at least
as important as ow stress characterisation of the work material. In many previously presented
machining analyses, the rake face contact stresses have been considered in terms of a con-
stant coe:cient of friction, , at the interface, based on the Coulomb friction: the relation be-
tween the friction and normal stresses, f and n (n positive in compression) has been simply
stated as
f = n : (3a)
While this rule represents good approximations under usual low contact stress conditions for
machine elements in sliding contact, it is now widely accepted as failing in the high contact
stress conditions on the rake face, near the cutting edge of a tool [16,27] and as reviewed
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2707

recently in Ref. [28]: above some critical normal stress, the friction stress becomes independent
of normal stress. Usui et al. [10] proposed the following friction model:


f = k 1 − e−(n =k) : (3b)
This expression reduces to Eq. (3a) at low values of n and saturates at the work material’s
local shear stress k at high values of n . Saturating at k is good for non-free-cutting steels, but
for free-cutting steels, a solid lubricant at the interface may reduce the saturation value to mk,
where 0 ¡ m ¡ 1. A modi3cation of Eq. (3b), to incorporate this, is


f = mk 1 − e−(n =mk) : (3c)
A further re3nement, which a6ects f in the transition between f = n and mk, is [28]


n 1=n
f = mk 1 − e−(n =mk) ; (3d)
where n is a constant.
Both simulations A and B have taken Eq. (3d) to represent sliding friction behaviour, but
their implementations of it have di6ered. In simulation A,  has been increased steadily, from
zero to its 3nal value, as the displacement between the work and tool has developed, in order
to ensure that the chip slides over the tool at all stages of the loading cycle. Simulation B has
supposed that there is initially no sliding between the chip and tool. Displacement between the
work and tool causes the contact shear stress to rise until it reaches the sliding friction value,
then sliding starts. In both simulations, once sliding has started, the rate of increase of friction
stress with normal stress, as work=tool displacement is increased step-by-step, has been found
from di6erentiation of Eq. (3d):

df f n n −(n =mk)n
= e : (4)
dn n f
Finally, the chip contact length is determined in the same way in both simulations. As
work=tool displacement increases, the nodes on the face of the chip that would contact the tool
are taken out of contact if a tensile normal contact stress is estimated at their position and are
brought into contact if their displacements cause them to overlap the tool.

3. Experimental studies: materials’ ow, thermal and friction behaviour

The simulations have been compared with experimental cutting tests, reported in Section 4.
Here the material property data input to the simulations are given. The cutting tests have
been carried out by turning 60 mm diameter hot rolled bars of three compositions. In order
of decreasing machinability, the 3rst one was a leaded low carbon re-sulphurised free-cutting
steel (PbLCFCS), the second was a plain low carbon re-sulphurised free-cutting steel (LCFCS)
and the third was a medium carbon re-sulphurised free-cutting steel (MCFCS). Their measured
compositions are given in Table 1.
Flow stresses of these steels were measured by incremental-strain split-Hopkinson bar rapid-
heating compression tests, over the equivalent strain range 0 –1, strain rate range 10−3 –103 and

at temperatures from room temperature to 700 C, following the method described in
2708 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

Table 1
Compositions of machined steels by wt%

Steel C Si Mn P S Cr Ni N Pb O

PbLCFCS 0.08 0.04 1.49 0.073 0.42 0.02 0.02 0.043 0.27 0.014
LCFCS 0.09 0.009 1.01 0.059 0.33 0.03 0.04 0.0036 ¡ 0:005 0.012
MCFCS 0.44 0.13 1.45 0.013 0.30 0.15 0.022 0.008 — —

◦ ◦
Fig. 6. Flow stresses at an equivalent strain of 1 and strain rate of 1000=s, measured from 20 C to 700 C and


extrapolated to 1000 C by Eqs. 1 or 2, for LCFCS ( ), PbLCFCS () and MCFCS ( ).

Ref. [14]. Fig. 6 shows the thermal softening observed at an equivalent strain of 1 and strain
rate of 1000: in e6ect it shows the variation of B with temperature. The parameters of the full

descriptions, following Eqs. (1) and (2), in units of MPa and C, are given in Table 2.
The work and tool thermal properties have been obtained from a handbook [29] and manufac-
turer’s information, respectively. They are listed in Table 3. Simulation A used room temperature
thermal data simulation while B included work variation with temperature.
Friction behaviour on the rake face was investigated by split-tool tests [16,28] in the same
conditions as the main cutting tests (Section 4): cutting speed from 50 to 250 m=min, at a
feed of 0:1 mm=rev, with a zero rake angle tool and without any uid lubrication=cooling. The
measured dependence of f on n was 3tted to the form of Eq. (3d). The results are presented
in Table 4.
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2709

Table 2
Work material ow stress parameters in Eqs. (1) and (2)

Material Parameters B (in MPa), M; N; k ∗ ; m∗ , and T (in C)
2 2
B = 960e−0:0012T + 210e−0:00005(T −290) + 100e−0:00005(T −590)
2
PbLCFCS N = 0:15e−0:0012T + 0:1e−0:00003(T −330)
M = 0:000018T + 0:023 k ∗ = 0:00025 m∗ = − 0:0021
2 2
B = 930e−0:0011T + 120e−0:00004(T −280) + 50e−0:0001(T −600)
2
LCFCS N = 0:16e−0:0017T + 0:09e−0:00003(T −370)
M = − 0:000038T + 0:018 k ∗ = 0:00025 m∗ = 0:0026
2 2
B = 1330e−0:0016T + 210e−0:00012(T −250) + 260e−0:00001(T −400)
2
MCFCS N = 0:11e−0:0016T + 0:09e−0:000035(T −475)
M = 0:000018T + 0:019 k ∗ = 0:00009 m∗ = − 0:0026

Table 3

Work and tool thermal conductivities and heat capacities (T in C)

Material K (W=m C) Cp (MJ=m3◦ C)

PbLCFCS 62 − 0:044T 3:53 + 0:003T


LCFCS 62 − 0:044T 3:53 + 0:003T
MCFCS 55 − 0:034T 3:53 + 0:003T
Toola 46 Not needed
a
A P20 cemented carbide: wt% WC:72, TiC=TaC:22.5, Co:5.5.

Table 4
Friction constants from split-tool tests

Cutting speed PbLCFCS LCFCS MCFCS


(m=min)  m n  m n  m n

50 0.75 0.05 1.3 0.80 0.75 2.2 1.5 0.70 1.0


100 0.77 0.52 1.3 0.90 0.78 2.2 1.6 0.68 1.0
150 0.80 0.54 2.2 1.0 0.80 1.7 1.7 0.66 1.0
200 1.05 0.64 1.8 1.3 0.80 1.7 1.6 0.67 1.0
250 1.3 0.74 1.8 1.6 0.80 1.7 1.5 0.66 1.0

4. Main cutting tests and comparison with simulations

Orthogonal turning tests were carried out at cutting speeds of 50, 100, 150, 200 and 250 m=min,
at a feed of 0:1 mm=rev, a depth of cut of 2:5 mm, a zero rake angle tool, in dry condi-
tions. Cutting and thrust forces were measured with a Kistler force dynamometer. The temper-
ature distributions along the rake face were directly measured with a single wire thermocouple
2710 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

Fig. 7. Comparison of measured and simulated cutting and thrust forces per unit depth of cut and mean contact
temperatures, for the turning of PbLCFCS steel.

Fig. 8. Comparison of measured and simulated cutting and thrust forces per unit depth of cut and mean contact
temperatures, for the turning of LCFCS steel.

embedded in the tool: the wire, made of tungsten, 20 m diameter, made contact with the chip
material in the contact area and the emf generated was calibrated as described in Ref. [30].
Figs. 7–9 show the forces and average rake face contact temperatures measured for the three
materials, and compares them with the simulation A and B predictions.

5. Discussion

Both simulation methods give good agreement with the experimental observations as far as
cutting forces are concerned. But simulation A has a tendency to underestimate the measured
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2711

Fig. 9. Comparison of measured and simulated cutting and thrust forces per unit depth of cut and mean contact
temperatures, for the turning of MCFCS steel.

average rake face temperatures (while simulation B gives good agreements), and simulation B
has a tendency to underestimate the thrust forces (while simulation A is better in this respect).
Subsidiary simulations, in which the dependencies of thermal conductivity and heat capacity on
temperature have been removed from simulation B, suggest that its good temperature predictions
depend on the inclusion of temperature-dependent work thermal properties. As far as thrust
forces are concerned, simulation B tends to underestimate the chip=tool contact length, whereas
A does not. The most likely cause of this is the di6erence between the two simulations in
applying the chip=tool friction boundary condition: namely (as described in Section 2) that A
allows the chip to slide over the tool from the very start of loading, whereas B does not.
The aim of this paper is to establish how robust is the ICM method in predicting steady-state
conditions. The di6erent loading paths associated with the di6erent applications of the fric-
tion boundary condition is the only aspect that has been found to in uence the 3nal re-
sult. The di6erent meshes (provided they are both made 3ne enough in the primary and
secondary shear regions), the di6erent approaches to separation of the chip from the work-
piece at the tool cutting edge and the di6erent convergence criteria have not in uenced the
results.
The establishment of realistic work material ow stress and rake face friction data remains
as the main barrier to accurate simulation of steady-state chip formation. In the present ex-
periments, the predicted temperatures have hardly risen above the temperature range in which
workpiece ow stress data have been gathered, but in many instances, extrapolations of ow
stress data have to be made [26]. As far as friction is concerned, in identical experiments to
◦ ◦
those reported here, but with a tool rake angle of 5 instead of 0 , simulations using the fric-

tion data gathered at a rake angle of 0 did not give such good agreement with experiment as
reported here [31]. The di6erent temperatures generated led, it is believed, to di6erent  and
m values at a given cutting speed. Linking friction parameters to temperature is considered in
Ref. [32].
2712 M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713

6. Conclusions

Two variants of the iterative convergence method (ICM) for elastic–plastic, coupled thermal,
3nite element simulation of steady-state chip formation in metal cutting have been tested against
experiments to assess what aspects are most important for accurate predictions. Allowing work-
piece thermal conductivity and heat capacity to vary with temperature, and the treatment of
sliding at the chip=tool interface, have emerged as the main important di6erences between the
simulations, once a 3ne enough mesh to deal with rates of spatial change in the primary and
secondary shear regions has been chosen. Both simulations have used the same material and
friction property data: obtaining such data remains the main issue for accurate simulation of the
machining process.

Acknowledgements

This work was partially supported by British Steel (now Corus Group) plc and the Euro-
pean Coal and Steel Community (ECSC). The authors are grateful to Professor Kitagawa of
Kitami Institute of Technology, Japan, for carrying out some of the experiments. The work
was undertaken while KM was a visiting fellow at the University of Leeds, funded by the
Japanese Ministry of Education and the UK Engineering and Physical Sciences Research Council
(EPSRC).

References

[1] Ernst H, Merchant ME. Chip formation, friction and high quality machined surfaces. In: Surface treatment of
metals. New York: ASM, 1941. p. 299 –378.
[2] Lee EH, Sha6er BW. The theory of plasticity applied to a problem of machining. Journal of Applied Mechanics
1951;18:405–13.
[3] Hasting WF, Mathew P, Oxley PLB. A machining theory for predicting chip geometry, cutting forces etc.
from work material properties and cutting conditions. Proceedings of the Royal Society of London 1980;
A-371:569–87.
[4] Childs THC. Elastic e6ects in metal cutting chip formation. International Journal of Mechanical Sciences
1980;22:457–66.
[5] Klamecki BE. Incipient chip formation in metal cutting—a 3-D 3nite element analysis. Ph.D. dissertation,
University of Illinois at Urbana Champaign, 1973.
[6] Strenkowski JS, Carroll JT. A 3nite element model of orthogonal metal cutting. Journal of Engineering and
Industry, ASME 1985;107:349–54.
[7] Iwata K, Osakada K, Terasaka Y. Process modelling of orthogonal cutting by rigid-plastic 3nite element
method. Journal of Engineering and Industry, ASME 1984;106B:132–8.
[8] Dudzinski D, Molinari A. A modelling of cutting for viscoplastic materials. International Journal of Mechanical
Sciences 1997;39:369–89.
[9] Usui E, Shirakashi T. Mechanics of machining—from descriptive to predictive theory. ASME Publ., PED
Vol-7, 1982. p. 13–35.
[10] Usui E, Maekawa K, Shirakashi T. Simulation analysis of built-up edge formation in machining of low carbon
steels. Bulletin of the Japanese Society of Precision Engineering 1981;15:237–42.
[11] Childs THC, Maekawa K. Computer aided simulation and experimental studies of chip ow and tool wear in
the turning of low alloy steels by cemented carbide tools. Wear 1990;139:235–50.
M.H. Dirikolu et al. / International Journal of Mechanical Sciences 43 (2001) 2699–2713 2713

[12] Childs THC, Otieno AMW, Maekawa K. The in uence of material ow properties on the machining of steels.
In: Proceedings of the Third International Conference on The Behaviour of Materials in Machining, Institute
of Materials, London, 1992. p. 104 –19.
[13] Childs THC, Dirikolu MH, Sammons MDS, Maekawa K, Kitagawa T. Experiments on and 3nite element
modelling of turning free-cutting steels at cutting speeds up to 250 m=min. In: Proceedings of the 1st French
and German Conference on High Speed Machining (HSM), Universite de Metz, Metz, 1997. p. 325 –31.
[14] Shirakashi T, Maekawa K, Usui E. Flow stress of low carbon steel at high temperature and strain rate
(Parts I–II). Bulletin of the Japan Society of Precision Engineering 1983;17:161–72.
[15] Gordon MB. The applicability of the binomial law to the process of friction in the cutting of metals. Wear
1967;10:274–90.
[16] Kato S, Yamaguchi K, Yamada M. Stress distribution at the interface between tool and chip in machining.
Journal of Engineering and Industry, Transactions of the ASME 1972;94B:683–9.
[17] Merchant ME. Mechanics of the metal cutting process II. Plasticity conditions in orthogonal cutting. Journal
of Applied Physics 1945;16:318–24.
[18] Shaw MC, Finnie I. The shear stress in metal cutting. Transactions of the ASME 1955;February:115 –25.
[19] Sata T. Recent developments concerning cutting mechanics. International Research Production and Engineering
ASME, NY 1963. p. 18–25.
[20] Nagtegaal JC, Parks DM, Rice JR. On numerically accurate 3nite element solutions in the fully plastic range.
Computer Methods in Applied Mechanics and Engineering 1977;4:153–77.
[21] Maekawa K, Childs THC. Finite element formulation of the Nagtegaal–Rice functional using constant strain
triangles. Journal of Faculty of Engineering of Ibaraki University 1992;39:53–66.
[22] Childs THC, Maekawa K, Maulik P. E6ects of coolant on temperature distribution in metal machining.
Materials Science and Technology 1988;4:1006–19.
[23] Mori K, Osakada K, Fukuda M. Simulation of severe plastic deformation by 3nite element method with
spatially 3xed elements. International Journal of Mechanical Sciences 1983;25:775–83.
[24] Becker ER, Carey GF, Oden JT. Finite elements, an introduction, Vol. 1. New Jersey: Prentice-Hall, 1981
[Chapter 5].
[25] Rowe GW, Sturgess CEN, Hartley P, Pilinger I. Finite element plasticity and metal forming analysis.
Cambridge: Cambridge University Press, 1991.
[26] Childs THC, Maekawa K, Obikawa T, Yamane Y. Metal machining: theory and applications. London: Arnold,
2000 [Chapter 6.4].
[27] Zorev NN. Interrelationship between shear processes occurring along tool face and on shear plane in metal
cutting. In: Proceedings of the Conference on International Research in Production Engineering, ASME,
New York, 1963. p. 42–9.
[28] Childs THC, Maekawa K, Obikawa T, Yamane Y. Metal machining: theory and applications. London: Arnold,
2000 [Chapter 2.1].
[29] ASM Metals Handbook. 10th ed., Vol. 1. Metals Park, OH: ASM, 1990.
[30] Childs THC, Maekawa K, Obikawa T, Yamane Y. Metal machining: theory and applications. London: Arnold,
2000 [Chapter 5.3].
[31] Childs THC, Bezas K, Dirikolu MH. The sensitivity of 3nite element predicted machining parameters to small
variations in ow stress behaviour. International Journal of Forming Processes 2000;3:99–114.
[32] Childs THC, Dirikolu MH, Maekawa K. Modelling of friction in the simulation of metal machining. In: 24th
Leeds–Lyon Symposium on Tribology. Amsterdam: Elsevier Tribology Series No. 34, 1998. p. 337– 46.

You might also like