You are on page 1of 15

SPECIAL ISSUE PAPER

197

Developments in modelling of metal cutting processes


S L Soo and D K Aspinwall
Department of Mechanical and Manufacturing Engineering, School of Engineering, The University of Birmingham,
Edgbaston, Birmingham, UK
The manuscript was received on 25 May 2007 and was accepted after revision for publication on 18 July 2007.
DOI: 10.1243/14644207JMDA163

Abstract: Following a brief historical perspective on the development of orthogonal cutting


models, including key work by Merchant and Oxley, the paper concentrates on the use of finite
element techniques to simulate two-dimensional orthogonal turning and the subsequent transition to three-dimensional formulations, thus enabling milling and drilling to be realistically
modelled. Input requirements including material models, mechanical property data at elevated
strain rates and temperatures, contact conditions, and other boundary conditions needed to
ensure accurate simulation and predictions are discussed. Current academic research and commercial developments involving simulation of non-conventional processes, microscale cutting,
workpiece surface integrity, and microstructure modelling, etc., are also highlighted.
Keywords: finite element modelling, machining, three dimensional

INTRODUCTION

The development of models to predict physical


behaviour during metal cutting, has been a major
preoccupation within the machining research community over the past 60 years. From an economic
viewpoint, it is evident that the ability to forecast key
process parameters such as tool life, cutting forces,
component surface roughness, and workpiece surface integrity, would be highly desirable as a means
of realizing cost savings, increased productivity, and
efficiency. Much of the early work revolved around
analytical models that represented the basic physics
and mechanics of metal cutting, using mathematical
relationships. With these equations, the shear angle,
tool/workpiece stresses, cutting forces, and temperatures can be estimated. However, all analytical models
are based on simplifying assumptions that do not
necessarily hold true under certain conditions. Nevertheless, despite such inherent weaknesses, research
has continued to the present day in an effort to create
more reliable and accurate models.

Corresponding

author: Department of Mechanical and Man-

ufacturing Engineering, School of Engineering, The University


of Birmingham, Edgbaston, Birmingham B15 2TT, UK. email:
s.l.soo@bham.ac.uk
JMDA163 IMechE 2007

The advent of computer-based simulation and in


particular development of the finite element (FE)
method, arguably brought about a step change in
the modelling of machining processes. First used by
Tay et al. [1] in the early 1970s and subsequently by
many other manufacturing researchers worldwide, the
technique has proven to be a powerful and comprehensive tool for studying the metal cutting process.
Among the many advantages offered by FE simulation over traditional mechanistic models is the ability
to accommodate the non-linear nature of the process
within the analysis. It is possible to undertake simulation of chip morphologies, evaluation of the effects
of tool geometry (e.g. nose radius, etc.) and analysis of
process variables with output measures such as cutting forces, temperature, and workpiece/tool stresses,
over the duration of the cutting operation.
2

ORTHOGONAL CUTTING MODEL

Orthogonal cutting is a term that was coined to


describe the case where the cutting edge of the tool
is straight and normal to both the cutting and feed
direction. Piispanen [2], was the first to introduce
a model to illustrate the orthogonal cutting process
in the late 1930s. His so-called card model involving blockwise slip depicted the workpiece material
being cut as a stack of cards, where thin lamellas
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

198

S L Soo and D K Aspinwall

angle and friction angle by equation (1)


=

Fig. 1

Piispanens card
formation [3]

stack

model

of

chip

are produced which slip against successive elements


as the tool penetrates the workpiece (Fig. 1) [3]. This
model though crude, found acceptance because of its
simplistic and understandable nature. It formed the
basis for Merchant to develop what is today widely
known as the idealized orthogonal cutting model
for describing the basic mechanics of chip formation
(Fig. 2) [4].

2.1

Merchants orthogonal force model

The well-known force model introduced by Merchant


was based on the orthogonal cutting configuration
and minimum energy principle, which states that the
shear angle will assume a value such that the total
work done in cutting will be a minimum [5]. Another
assumption is that the shear strength of the material
being machined is independent of the shear angle ,
but equivalent to the shear stress acting on the shear
plane. Based on the above suppositions, Merchant
showed that the shear angle is related to the tool rake

Fig. 2 Idealized orthogonal cutting model [4]


Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

1
( )
4
2

(1)

In his model, Merchant considered the chip as a separate body in equilibrium under the influence of two
opposing equal resultant forces denoted as R and R  ,
respectively. R is the force exerted by the tool on the
back surface of the chip while R  is the force exerted by
the workpiece acting on the shear plane [6]. This force
system is shown in Fig. 3 [6], where R  is resolved into
a component F parallel to the tool face, which is the
friction force acting when the chip slides over the tool,
and N , which is the normal force acting perpendicular
to F . The friction angle is simply the angle between
the R  and N .
In the same way, force R may be resolved into a
horizontal component Ft which is the tangential or
cutting force and a vertical component Ff which is the
feed force. Additionally, force R can also be resolved
into a force vector along the shear plane known as the
shearing force FS , which is responsible for the work
expended during shearing of the material, and into a
component FN normal to FS , which exerts a compressive stress on the shear plane [7]. Given that R and R 
are equal and parallel (but not necessarily collinear),
all their associated components can be represented by
a single resultant force vector R located at the tool tip
for convenience. This arrangement is shown Fig. 4 [6]
and is popularly referred to as the Merchants circle
force diagram or simply Merchants circle [8].
By considering the geometry of the system together
with several simplifying assumptions, forces acting on
the shear plane and tool chip interface and subsequently stresses in the primary and secondary shear
zones can be determined/calculated. Unfortunately,

Fig. 3

Force system diagram [6]


JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

Fig. 5

Fig. 4

Merchants circle force diagram [6]

when applied to a range of workpiece materials, the


model does not provide accurate results, indeed it does
not satisfactorily account for the influence of parameters such as cutting speed on machining behaviour [8].
The description of friction at the tool/chip interface
using the Coulomb model is also thought to contribute
to the weakness of the analysis. Nevertheless, Merchants circle remains an important milestone in metal
cutting theory as it provides fundamental insight into
the forces and stresses involved in machining.
2.2

199

Chip formation occurring within a finite plastic


zone [11]

AB are sliplines that represent the directions of maximum shear stress and shear strain rate. The work
velocity is assumed to change to the chip velocity
in the shear zone without any discontinuities, however, the resultant force R does not in general pass
through the midpoint of AB. Furthermore, chip curl is
neglected, while the state of strain and consequently
the shear flow stress along each of the parallel sliplines
is constant [11]. The method of analysis involved
determining the stresses along AB as a function of the
shear angle and work material properties, and subsequently selecting a value for such that the resultant
force transmitted by AB is in equilibrium with the
frictional conditions at the tool-chip interface (consistent with the direction of ) [12]. Once the shear
angle is found, various other components of force
can be determined. A full description and treatment
of the theory can be found in reference [12]. Despite

Oxleys parallel sided shear zone theory

In re-examining Merchants theory with regard to


shear angle prediction, Oxley found that when
a wide range of cutting conditions was applied,
Merchants theory showed very poor quantitative
agreement with experimentally measured values [9].
Oxley subsequently criticized some of the necessary
assumptions, not least the disregard of effects due to
strain hardening, strain rate etc. [10]. Oxley and Welsh
[11] introduced a new predictive analytical model to
describe the orthogonal cutting process which took
account of such aspects; furthermore, direct observations of chip formation indicated that shearing
actually occurred within a finite plastic zone instead
of on a simple shear plane (Fig. 5).
The Oxley and Welsh model represented this finite
plastic zone in the form of a parallel-sided shear zone
as shown in Fig. 6. Here the shear plane AB is assumed
to extend to the boundaries marked by CD and EF,
respectively, which are parallel and equidistant to AB
and form the shear zone. The lines CD and EF like
JMDA163 IMechE 2007

Fig. 6

Parallel sided shear zone model [11]

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

200

S L Soo and D K Aspinwall

the obvious improvements in Oxley and Welshs model


over some of its predecessors, the lack of sufficient
stress/strain data at the level of strain rates and temperatures encountered in machining is a significant
limitation.
In addition to the shear angle based equations such
as those developed by Merchant and Oxley for determining forces, stresses, strain, etc., there has also
been numerous theoretical models developed over the
years for the prediction of cutting temperatures at the
tool-chip interface and shear plane, such as those by
Trigger and Chao [13], Boothroyd [14], and Hou and
Komanduri [15] to name but a few. Successive analytical models have improved upon their predecessors
and given tremendous insight into the mechanics of
the chip formation process. Despite the advantages
offered by such solutions, analytical models have limited practical application while many of the respective
underlying assumptions do not hold true for a great
many materials and cutting conditions.
3

FE MODELLING

In the two decades following Tay et al.s [1] pioneering work, academic research continued to centre
mainly on two-dimensional, orthogonal-based models with workpiece materials ranging from low carbon steel, copper, stainless, and hardened steels to
more exotic alloys such as titanium and nickel based
superalloys (Inconel 718). The following sections discuss the key requirements for realistic FE simulation of metal cutting, together with the more recent
thrust towards more complex/representative threedimensional-based models, including processes other
than turning. Finally, recent developments which have
focused on areas such as surface integrity prediction
(residual stress, microstructure, etc.) after machining
as well as simulation of microscale mechanical cutting and non-traditional machining processes such
electrical discharge machining (EDM) are presented.
3.1

Model formulation

This refers to the way in which the FE mesh is associated with the workpiece material. The three main
formulations are the Eulerian, Lagrangian, and the
arbitrary LagrangianEulerian (ALE).
3.1.1

Eulerian

In the Eulerian representation, the FE grid is spatially


fixed, with the material flowing through the meshed
control volume [16]. The unknown material variables
are calculated at set locations as the material flows
through the mesh. Eulerian based models are free from
element distortion problems as their shapes do not
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

alter throughout the course of the simulation, and


hence no remeshing is necessary, which is a key benefit in terms of computational efficiency. Furthermore,
a chip separation criterion to designate the parting of
nodes between the workpiece and chip is not required,
as the material is not attached to the mesh. The nature
of Eulerian type formulations, however, means that the
initial shape of chip and contact conditions must be
known or assumed in advance. This then excludes the
possibility of modelling unconstrained flow of material at the free boundaries and the natural formation
of the chip as the result of deformation. The free surfaces of the chip must therefore be adjusted manually
during a simulation, usually through an iterative procedure [17]. In general, this involves checking the
velocities at the nodes along the free boundaries to
ensure that their components, which are normal to
it, are zero and that other variables have converged
[18]. The geometry of the chip is then updated with the
computed velocities tangential to the free surfaces.
The method is suitable for studying steady-state
cutting, where the relatively lengthy transition from
transient incipient chip formation to steady-state conditions are not of interest [19]. Strenkowski and Moon
[20] utilized an Eulerian FE model to predict steadystate chip geometry and chip/tool contact length. The
latter parameter was determined by inspecting the
normal stress at each node along the tool-chip interface. A positive magnitude meant that the particular
node had been released from the interface. Although
reasonable predictions have been reported by several
investigators using Eulerian based models [21, 22], it
is widely agreed to be better suited for fluid mechanics
problems rather than for machining.
3.1.2

Lagrangian

Here the FE mesh is attached to the workpiece material


and the elements deform together with the material during cutting. This makes it highly suitable for
solid mechanics analysis and appropriate for problems where unconstrained material flow is involved
[18]. For metal cutting simulations, the Lagrangian
formulation is preferable due to the more convenient
modelling of the evolution of the chip from the incipient stage to steady. The geometry of the material
boundaries (or chip shape) does not have to be predetermined, but develops during the course of the
analysis entirely as a function of the physical deformation process, machining parameters, and material
properties [23].
The principal disadvantage of the Lagrangian
approach in machining simulations is that as the elements generally experience severe distortion, geometrical as well as material non-linearities are introduced
in the FE equations [17] that considerably increase the
computational load. This poses substantial numerical
JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

difficulties, which occasionally requires the mesh to


be regenerated to prevent the simulation from breaking down prematurely, besides adversely impacting
on the efficiency and accuracy of the analysis. In
Lagrangian based models, the parting of the chip from
the workpiece is traditionally achieved via separation
of nodes in front of the tool tip along a predefined
line representing the depth of cut [2427]. The node
separation procedure is governed by a material failure condition, which is a function of one of several
criteria and although simple and straightforward, the
methodology does possess drawbacks. With the large
deformations encountered during metal cutting, there
is a tendency for the nodes in front of the tool along the
parting line to be pushed out of position, thus causing entanglement of the elements with the cutting
tool. Premature separation of nodes resulting in a gap
in front of the tool tip is another common problem,
which is usually caused by inappropriate specification of the magnitude of the separation criterion.
Such effects contribute to reducing the accuracy and
validity of the results. More recently, however, several
researchers have investigated alternative methods of
achieving the cutting action such as element deletion
[28, 29] and adaptive remeshing [30].
3.1.3

Arbitrary Lagrangian Eulerian (ALE)

The ALE procedure can be described as a general formulation that amalgamates both the classical
Lagrangian and Eulerian techniques into one description in order to exploit their respective merits. The FE
mesh in an ALE simulation is neither fixed spatially
nor attached to the material, but instead is allowed to
move arbitrarily relative to the material [31]. The formulation is such that it can be reduced to Eulerian
or Lagrangian descriptions as and when required [32].
For metal cutting simulations, the idea is generally to
apply features of the Eulerian type approach for modelling the area around the tool tip, while the Lagrangian
form can be utilized for modelling the unconstrained
flow of material at the free boundaries [33]. In this
way, the problem of severe element distortion and
entanglement in the cutting zone can be alleviated
without the need for remeshing. Furthermore the evolution of the shape and size of the chip can occur
freely and automatically as a function of the material
deformation. A number of researchers have studied
the problem of orthogonal turning with continuous
chips using the ALE procedure [34, 35].
3.2 Workpiece material constitutive
equations/material properties testing
Constitutive equations describe the mechanical
behaviour of the workpiece material undergoing loading and deformation, and are an essential input in any
JMDA163 IMechE 2007

201

FEM model. Under machining conditions, the workpiece material is generally subjected to high levels of
heat, strain and strain rate, which significantly influences flow stress. In FE simulations, an ideal plasticity
model for the workpiece would be one that can reliably describe the stressstrain response together with
its dependence on strain-rate, temperature and work
hardening [36]. The material property data for constitutive modelling was previously acquired through
standard mechanical testing such as tension, compression, torsion or impact experiments. However,
these test methods normally cannot attain the level of
strains and strain rates that are encountered in typical
machining processes [37]. Assumptions are therefore customarily introduced according to the need of
the application, partly to overcome the inadequacies
imposed by the experimentally obtained material data
and to simplify the modelling [36]. A brief discussion of
two of the general constitutive plasticity models with
strain rate and temperature dependence that are available in the commercial FE software ABAQUS (widely
used for metal cutting simulations) is presented.
The level of plastic deformation is such in the primary and secondary shear zones that many modellers
discount the elastic response of the workpiece material in their simulations for the sake of expediency.
Nonetheless, elastic straining although small (usually
of the order 103 ), does influence aspects such as
the residual stress and strain distribution in the chip
and workpiece. In ABAQUS, the elastic (recoverable)
and inelastic (non-recoverable) responses are separable based on an assumption that there is an additive
relationship between the strain rate [38], such that
= el + pl

(2)

where is the total strain rate, el the rate of change


of elastic strain, and pl the rate of change of plastic
strain.
The elastic response is modelled by the relatively
simple linear elasticity relationship given by
= Del : el

(3)

where is the total stress, el the elastic strain, and


Del is the elasticity tensor matrix that may depend on
temperature.
The classical plasticity model available in ABAQUS
is based on the von-Mises yield surface used to define
isotropic yielding and hardening. Isotropic yielding
occurs when the yield stress increases in all directions
as plastic straining occurs. The yield function can be
written as
f ( ) = 0 ( pl , )

(4)

where 0 is the equivalent (uniaxial) stress, pl the


plastic strain, and the temperature.
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

202

S L Soo and D K Aspinwall

Strain rate effects are introduced into the constitutive material model through an overstress power law
relationship which is expressed by
pl = D

p

1
0

for

> 0

(5)

where pl is the equivalent plastic strain rate, is the


yield stress at non-zero strain rate, 0 is the static yield
stress (which may be a function of the plastic strain
and temperature), and D and p are material properties
that can also be functions of temperature.
Another plasticity model provided in ABAQUS that
has been widely used is the JohnsonCook constitutive
relationship. This was developed with data obtained
from a series of Hopkinson bar tests at high strain
rates and temperatures on copper, iron, and AISI 4340
steel [39]. The formulation also reflects a particular
type of von-Mises plasticity where the flow stress is
expressed as
 pl 

= [A + B( pl )n ] 1 + C ln
0


m 
transition
1
melt transition


(6)

where pl is the equivalent plastic strain, 0 is the initial


plastic strain rate (normally taken to be 1.0 s1 ), melt
is the melting temperature of the workpiece, transition
is the transition temperature defined as the one at
or below which there is no temperature dependence
(usually taken as room temperature), while A, B, C, n,
and m are material constants.
The JohnsonCook constitutive equation basically
considers the mechanical behaviour of materials as
multiplicative effects of strain, strain rate, and temperature. It is a relatively simple model to apply, as most
of the variables are readily adaptable to computer
codes [36].
As indicated previously, workpiece material strain
rates are known to exceed 104 s1 within the shear
zone, depending on the cutting parameters specified. The lack of appropriate high strain rate material
data is one of the stumbling blocks to achieving
accurate/realistic predictions from FE simulations of
metal cutting processes, which has been highlighted
by Childs [40]. Experimental techniques for obtaining
high strain rate data (ideally at elevated temperatures)
may utilize specialist commercial testing equipment
(tensile or compression testing) such as the Gleeble
system [41] albeit with achievable strain rates limited
to 1000/s depending on the workpiece. Alternatively,
the Split Hopkinson pressure bar device, details of
which can be found in reference [42], can be employed.
Here an explosive impact configuration is employed,
however the equipment is generally bespoke and not
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

commonly available. Irrespective of the test methods used, however, the costs and effort involved in
generating the required material property data at elevated strain rates and temperatures are considerable.
Although a set of material property data produced
using a given technique can be fitted to a number
of different constitutive equations, Shi and Liu [43]
showed in a recent paper that the choice of material model can significantly influence the predicted
results. Similarly, Umbrello et al. [44] suggested that
FE output results can also be sensitive to the workpiece
material constants specified for a particular material
model (in this case, the JohnsonCook flow stress relationship was utilized). They concluded that this was
due to the different test methods by which each set of
material constants were found/generated.
3.3

Chip separation criteria/chip segmentation


requirements

As outlined in section 3.1.2, the triggering of material fracture, and hence parting of the chip from
the workpiece in Lagrangian based models, is governed by a separation criterion. Most commercial
FE packages provide the capability for debonding or
separation of elements for general fracture mechanics or crack propagation studies, and have been
adopted by researchers for metal cutting simulations.
To a large extent, the software selected limits the
choice of separation criterion available to the modeller, unless custom algorithms/codes are written and
integrated into the model. An ideal criterion would
be one that is indicative of the physical mechanism
of chip formation for a material under machining
conditions.
Strenkowski and Carroll [27] applied an effective
plastic strain criterion to simulate orthogonal cutting
of a titanium alloy using the commercial NIKE2D software. The effective plastic strain at the node closest
to the tool tip was compared with a specified rupture
value for the workpiece material. Once the predicted
plastic strain exceeds the critical value for failure, the
node was released and allowed to move away from
the workpiece. This approach has also been employed
by a number of other researchers [45, 46]. Another
widely applied criterion is based on a distance tolerance [47] where chip separation is initiated when
the length between nodes along the parting line and
tool edge is less or equal to the magnitude of a user
predefined distance [25, 26]. One of the main challenges with the distance criterion is that it must be
extremely small (usually of the order of microns) in
order to maintain realistic results, but at the same time
preserve numerical stability.
Shet and Deng [24] specified a critical stress criterion using the commercial software ABAQUS to
activate node separation. This is somewhat similar to
JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

the effective plastic strain criterion, except that the


normal component of stress is taken as the failure
measure. The choice of chip separation criterion and
its corresponding magnitude is a source of controversy
among the modelling community. A critical look at
the literature suggests that the selection of criteria is
largely arbitrary in nature while experience and trial
and error have formed the basis for determining the
criterion value. For example, Zhang [48] argued that
although plastic strain is a quantity that can be related
and used to define material failure, its value changes
significantly during the transition from transient to
steady-state cutting, and also when the cutting parameters are varied. Therefore, a single threshold value of
effective plastic strain cannot be expected to act as a
reliable criterion.
Much of the early work in FE modelling of the cutting
process involved the simulation of continuous chips.
These models though useful, could not predict the
chip morphology and associated parameters of materials that exhibited segmented or discontinuous chips.
Material models available in commercial FE software
(such as ABAQUS) are designed for use across a wide
spectrum of applications, and are usually inadequate
to simulate chip segmentation on their own. To model
this phenomenon, a suitable damage criterion, analogous to that specified to initiate chip formation, is
necessary in order to numerically describe the onset
and propagation of shear localization followed by fracture during deformation of the chip. It was not until
the mid-1990s that the first models dealing with segmented/discontinuous chip formation appeared in
the literature [30, 49]. With commercial FE packages,
a customized user subroutine containing the damage
criterion must be written and incorporated into the
input FE code. Researchers who have previously been
successful in modelling two-dimensional segmented
chips have confirmed the need for these subroutines
[28, 50] and several of these are briefly reviewed.
Marusich and Ortiz [51] applied a fracture criterion
based on critical stress for brittle mode fracture and
another criterion based on effective plastic strain for
ductile failure. Brittle fracture was deemed to occur
when the hoop stress ( ) on a plane attained a critical value f , corresponding to the material fracture
toughness KIC given by
KIC
f =
2l

(7)

and
max (l, ) = f

(8)

Similarly, ductile fracture proceeded if the effective


p
plastic strain p reached a critical value f expressed by
p

max p (l, ) = f

JMDA163 IMechE 2007

(9)

203

with the crack propagating at the angle for which


the criterion is met, where l is the critical distance in
front of the crack tip. The critical plastic strain was then
estimated to be
f 2.48e 1.5p/
p

(10)

where p is the hydrostatic pressure.


Obikawa et al. [52] in simulating discontinuous chip
formation when machining brass, also applied a damage criterion based on equivalent plastic strain. The
fracture strain c in this case was a function of equivalent stress ( ), strain rate ( p ), cutting speed (Vc ), and
hydrostatic pressure (p) expressed by
c = 0

p
p

Vc

(11)

where 0 , , and are constants.


Crack nucleation and growth occurred when the
equivalent plastic strain exceeded the fracture strain.
Ceretti et al. [53] utilized the Cockroft and Latham
damage parameter to model chip segmentation with
the commercial FE software DEFORM2D. The damage
was evaluated according to the equation
 f
Ci =
d
(12)
0

where Ci is the critical damage value given by a uniaxial tensile test, f is the strain at fracture, is the
maximum stress and is the effective strain. The criterion predicted the onset of fracture when the value of
Ci was exceeded. Deformed elements that fulfilled the
damage value were deleted with the boundary of the
removed elements smoothed to reduce the loss of volume that inevitably takes place. Hua and Shivpuri [54]
applied an almost identical approach when simulating
the orthogonal turning of Ti-6Al-4V alloy.
3.4 Tool-chip interface friction
The tool-chip interface is characterized by plastic
deformation on the underside of the chip as well as
intense heat dissipation due to work done to overcome the resistance imposed on chip flow by friction.
Cutting forces, stress distributions and tool wear in
particular are greatly influenced by the frictional relationships in this region yet measuring and modelling
the friction accurately is complicated. Many of the
early FE simulations on metal cutting published in the
literature have assumed that contact in the secondary
shear zone behaves according to Coulombs or Amontons law, which states that the frictional sliding force
(F ) is proportional to the normal load applied (N ). The
ratio of F over N is known as the coefficient of friction
, which under lightly loaded slider conditions is independent of the normal load. For this to apply however,
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

204

S L Soo and D K Aspinwall

N must be below a certain critical value. As the normal


force increases, Coulombs equation no longer holds
true as the real area of contact between the chip and
tool rake face increases [3]. This is known as a highly
loaded slider state.
As mentioned previously, the friction at the toolchip interface during cutting is very complex, but is
widely believed to be an amalgamation of both the
highly loaded as well as lightly loaded regimes. Usui
and Takeyama [55] studied the distribution of the
shear stress (C ) and normal stress (C ) along the rake
face of a photoelastic cutting tool (epoxy resin) when
machining lead at low speeds. They found that C
remains constant over two-thirds the distance of the
tool-chip contact length and subsequently decreased
gradually to zero at which point the chip loses contact with the tool. Conversely, the maximum C at the
tool tip was significantly higher than the maximum
C and decreased exponentially to zero. A simplified
schematic representation of their results is shown in
Fig. 7 [3]. The length AD where the shear stress remains
constant is also known as the sticking or seized region,
and is characterized by a large normal load analogous
to a highly loaded slider. The area DC is known as the
sliding region where both the shear stress and normal stress decrease to zero. Various friction models
based on this observation have since been developed
to numerically describe the interaction at the tool-chip
interface and subsequently utilized in FE simulations
of the machining process. A number of these models have been reviewed in a recent paper by Ozel
[56], which showed that predictions were more accurate when using frictional models that incorporated
variable shear stresses and coefficient of friction.

Fig. 7

Normal and shear stress distribution on rake face


of tool [3]

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

3.5

Simulation of machining processes two


dimensional to three dimensional

Much of the literature on FEM in metal cutting up


to the early 1990s concerned two-dimensional, plane
strain descriptions, allowing comparison with orthogonal turning of conventional workpiece materials
such as low carbon, mild, or free machining steel.
Despite dramatic improvements in computational
power and FE software robustness/sophistication over
the past decade, some researchers still persist with
two-dimensional formulations, as the scope for study
still exists [5761]. Although the simulation of continuous chip formation and general prediction of stress
strain distributions within the chip was the thrust of
early investigations, the later models have been developed for investigating more specific aspects associated
with the cutting process. This has included among
other things, effects of varying tool geometry [25], evaluation of workpiece/chip temperatures [62, 63], as
well as modelling the effect of coatings on temperatures generated in the secondary shear zone [6466].
Similarly in the mid-1990s, research progressed into
prediction of different chip morphologies, notably
shear localized/discontinuous chips which occur with
many workpiece materials such as hardened steels
and titanium alloys as a result of heavy localized
plastic deformation leading to adiabatic shear bands.
Chip shape can influence cutting forces and temperatures, which can have knock-on effects on tool
life and surface quality/integrity. Ng [50] was one of
the first to successfully model the transition in chip
morphology from continuous to segmental due to
variations in cutting speed and workpiece hardness,
when turning AISI H13 using PCBN tooling. This was
made possible by the incorporation of a customized
user defined subroutine that contained a crack nucleation module based on effective plastic strain together
with the use of element deletion. Guo and Yen [67]
also demonstrated the simulation of discontinuous
chip formation when hard machining AISI 4340 steel.
Similarly, the characteristic of Ti-6Al-4V to exhibit
plastic instability during machining leading to serrated/segmental chips has also been the subject of
considerable FE modelling studies [6870]. Here Bker
et al. [69] utilized a slightly different approach whereby
they developed a special algorithm that triggered automatic remeshing as a result of heavy distortion when
a shear band occurred. This remeshing technique
changed the mesh topology by adding free nodes
where necessary in order to create the chip segment.
Of late, the emphasis has shifted towards threedimensional computations with much of the initial
work centring on turning processes. One of the earliest three-dimensional investigations was presented
by Maekawa et al. [71] who analysed the machinability of leaded CrMo and MnB steel based on
JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

strain rate and temperature-dependent plastic flow


material properties. Their model was used to predict
the stress and temperature distribution together with
the wear rate on the rake face of the cutting tool. A
more recent contribution by Li and Shih [72] looked at
modelling the effects of cutting parameters and tool
geometry on cutting forces, temperature, and chip
curl/morphology when turning commercially pure
titanium. This involved an updated Lagrangian FE formulation together with adaptive meshing techniques
and the application of a coupled thermal and mechanical analysis of the tool and workpiece deformation.
In general, the results of their three-dimensional
model were successfully validated against corresponding experimental data, although the tangential force
was typically under-estimated by up to 20 per cent
at the lower cutting speeds tested. The discrepancy
was mainly attributed to the simplified friction model
applied and limitations in the material property data.
The simulation of alternative cutting processes such
as twist drilling and milling in particular, have started
to gain prominence in the last 5 years (Figs 8 and 9).
For such operations, a three-dimensional model is not
only desirable but necessary in order to fully realize

Fig. 8 Three-dimensional FE simulation of twist drilling


using DEFORM3D (courtesy of SFT Corporation)

Fig. 10

205

Fig. 9 Three-dimensional FE simulation of face milling


using AdvantEdge (courtesy of ThirdWave
Systems Inc.)

the problem due to the complex geometry and rotary


motion of the cutter.
Three-dimensional processes that have been simulated using FE methods include the drilling of stainless
steel as a means to evaluate burr formation [73],
twist milling of AISI 4140 steel [74] and simulation
of high speed ball nose end milling of Inconel 718
using ABAQUS Explicit [75] (Fig. 10). By studying
the stress contours and workpiece edge deformation
predicted by the model, Guo and Dornfeld [73] proposed a mechanism for burr formation comprising
four stages. These were outlined as burr initiation,
burr development, pivoting point and finally burr
formation. Although the height and thickness of the
simulated burrs were not compared to experimental data, the characteristics of the burr in terms of
shape and cap formation were validated by results
from drilling of 304L stainless steel and plasticine.
With the twist milling simulation, Bacaria et al. [74]
applied the JohnsonCook damage criterion over the
entire workpiece, which allowed the natural formation of the chip as the tooth of the milling cutter
penetrated the workpiece, without the need for a predefined fracture line (which is virtually impossible

(a) Three-dimensional simulation of high speed ball nose end milling of Inconel 718 using
ABAQUS Explicit; (b) to (d) showing progression of chip formation [75]

JMDA163 IMechE 2007

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

206

S L Soo and D K Aspinwall

to specify due to the complex trajectory of milling


operations). The work, however, only dealt with preliminary chip formation aspects, together with some
stress distribution and cutting force predictions and
was not validated experimentally. A more recent contribution has been the three-dimensional modelling
of the friction drilling process involving an aluminium
alloy Al-6061-T6 using ABAQUS Explicit [76]. In contrast to conventional twist drilling, friction drilling
involves rotating a conical tool against a workpiece
material such that the resulting frictional contact
between the tool-workpiece generates sufficient heat
to soften the workpiece where the axial feed provides
penetration. The model by Miller and Shih [76] was
able to successfully simulate the material deformation
and temperature in the workpiece and was subsequently validated against experimental measurement
of torque, thrust force, and temperature. The torque
predictions in particular were found to deviate from
the experimental data at higher tool feed rates, which
was due to the limitation of the Coulomb friction
model.
3.6

Recent developments and future research

The use of FE modelling techniques as a design or optimization tool for bulk deformation processes such as
casting and to a lesser extent metal forming, is relatively accepted and widespread, particularly within
the aerospace industry. Unfortunately, this technology
is less developed with regard to machining processes
despite the progress detailed in the previous section.
Current commercial packages are not able to predict
with reasonable reliability and accuracy aspects such
as workpiece surface integrity postmachining, which
is of paramount importance with safety critical components such as aero-engine parts. There is, however,
a growing interest in developing suitable expertise
within this area, evident from the recent increase in
the number of publications on the subject, albeit with
two-dimensional formulations. Salio et al. [77] developed a two-dimensional plane strain model using
MSC MARC to evaluate the residual stress generated
after finish turning of Inconel 718 turbine disc material
over a wide range of cutting parameters. The predicted
results were validated against calculations from analytical models as well as experimental X-ray residual
stress measurements. In general, the modelled residual stresses were found to give good agreement when
compared with measured data. It was observed that
tensile residual stress values at the workpiece surface
increased with smaller depths of cut and the crossover
to a compressive state occurred at a deeper level into
the workpiece material. The work also identified preferred depth of cut values to provide minimum tensile
residual stresses.

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

Another related area that has received significant


attention is the effect of tool geometry (in particular tool edge radius) and associated cutting parameters in relation to workpiece residual stress. Hua
et al. [78] studied the effects of feed rate, workpiece
hardness, and tool geometry on the subsurface residual stress when hard facing bearing steel AISI 52100
using a Lagrangian FE approach with DEFORM2D.
Their simulations showed that in all conditions tested,
compressive residual stress was produced at the workpiece surface. An increase in feed rate and workpiece
hardness was found to significantly raise the magnitude of compressive stress generated, although the
latter did not alter the depth of maximum residual
stress. Similarly, an increase in tool hone radius and
chamfer angle gave rise to a higher maximum residual stress, with the former parameter having a more
dominant effect. Cutting speed was observed to have
no major impact on the residual stress response. The
predicted results were shown to agree closely with corresponding X-ray diffraction measurements of actual
machined surfaces. Nasr et al. [79] developed an ALE
model in ABAQUS Explicit to investigate the effect of
varying tool edge radius on subsurface residual stress
when orthogonal cutting AISI 316L stainless steel.
They found that an increased tool edge radius gave
rise to higher tensile residual stresses at the machined
surface but also induced larger compressive stresses
deeper into the workpiece. This was attributed to the
higher cutting temperatures generated with a larger
tool edge radius (wider area of contact) and the greater
plastic deformation as a result of increased material
being ploughed into the machined surface. Similar
methodologies to predict surface integrity have also
been utilized by Ee et al. [80] and Zong et al. [81] when
evaluating the orthogonal cutting of AISI 1045 steel
and diamond turning of OFHC copper, respectively.
The work by Ee et al. [80] also investigated the effect
of sequential/subsequent cuts on the residual stress
profiles and found that although no significant change
was recorded in terms of the maximum tensile stress
at the workpiece surface, the residual stresses decayed
to a neutral state at half the depth compared to the
case where only a single cut was made.
In addition to residual stress, the effects of machining processes on workpiece/component microstructure is likewise of great interest as they can have a
significant influence on part performance, properties,
and fatigue life. Chuzhoy et al. [82] presented one
of the earliest studies on modelling the machining
of ductile iron which has a heterogeneous structure
comprising pearlite and ferrite together with specified
amounts of graphite. The microstructure of the material was successfully generated and incorporated into
the FE mesh with the distribution and average size of
the grains chosen to be representative of the actual

JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

material. Each of the phases was explicitly modelled


with the hexagonal shaped pearlite and ferrite grains
interspersed throughout the material while the circular graphite nodules were embedded within the ferrite
particles. Furthermore, the corresponding properties
of each phase was individually characterized prior to
the cutting simulation and subsequently accounted
for in the material model. The orthogonal cutting simulation carried out on the multi-phase material was
capable of predicting cutting forces, temperatures,
stress/strain distributions together with damage and
changes within each phase. The technique was able to
illustrate the effect of the material microstructure on
machinability with the potential for alternative applications. More recent studies relating to microstructure
modelling when machining have included work by
Simoneau et al. [83] on chip formation in orthogonal turning of AISI 1045 steel and Pramanik et al. [84]
when investigating the tool-particle interaction in
cutting of SiC reinforced aluminium metal matrix
composites.
An important class of heterogeneous materials
which is currently receiving much attention in industry are fibre reinforced composite materials. As the
cost of composites is inherently expensive, the ability to investigate the cutting of such materials through
simulation rather than experimentation alone is very
attractive, indeed advances in FE techniques have led
to numerous publications on the subject of modelling
composites when orthogonal machining. A review
of the literature indicates that there are two principal methods to model the properties of composites,
namely use of an equivalent homogenous material (EHM) technique and a micromechanics based
approach. With the former, the material is defined
as a single homogenous structure with the properties of the fibre and matrix combined whereas the
micromechanics technique considers the fibre and
matrix properties separately within the mesh. This
enables modelling of local material behaviour effects
during machining but imposes a high computational
load while the EHM methodology is less computationally intensive, but is unable to simulate interactions
between the fibre and matrix. Arola and Ramulu [85]
published one of the earliest modelling studies on
orthogonal cutting of fibre reinforced composites utilizing the EHM approach. They reported that while the
predicted main cutting forces agreed well with experimentally measured results, the thrust forces were
inconsistent. Other works based on the EHM method
have also reported similar findings [86, 87]. More
recent research using micromechanics modelling has
managed to overcome the discrepancy in the thrust
force prediction and is able to model aspects such as
matrix damage and fibre failure [88]. Multi-scale modelling techniques integrating micromechanics theory
has also been used to simulate the laser assisted
JMDA163 IMechE 2007

207

cutting of silicon nitride ceramics [89]. Here, six-node


hexagonal continuum elements were used to represent the bulk workpiece, while thin interfacial cohesive
elements were applied around the bulk material to
enable initiation and crack propagation during the
cutting process. The model was successful in predicting discontinuous chip morphology and was validated
against experimental measurement of cutting force
and residual stress.
In recent years, interest in microscale mechanical machining has quickened with processes such as
high speed micromilling offering a viable and attractive manufacturing route for producing miniaturized
components for use in the bio-medical, telecommunications, automotive, and aerospace sectors [90].
Although still in its infancy, FE research on micromachining to complement experimental based research
has already begun to appear in the literature [91, 92].
Alternative modelling techniques such as molecular
dynamics simulations have also been introduced, in
particular for applications relating to nanometric level
cutting [92]. Other recent developments have been
in the FE simulation of non-traditional cutting techniques such as EDM [93] and electro-chemical spark
machining (ECSM) [94].

COMMERCIAL/INDUSTRIAL UTILIZATION

Up until the mid-1990s, all work involved either


custom made (usually written and tailored for a
specific process/problem) or general purpose commercial packages including ABAQUS, NIKE2D, ANSYS,
MARC, etc., and were almost exclusively the domain
of academia, as it required highly skilled specialists
for the formulation of models and interpretation
of results. In recent years, however, a commercial FE package AdvantEdge (by ThirdWave Systems
Inc.), tailored specifically for modelling metal cutting
processes has been available, with a similar product (DEFORM2D/3D) also offered through Scientific
Forming Technology (SFT) Corporation. Both systems incorporate highly user friendly modules such
as extensive parametric tool geometry and material
model libraries together with powerful adaptive meshing capabilities. The automatic/adaptive remeshing
procedure present in both programs allows chip
formation to progress without the traditional material separation mechanisms such as node separation
or element deletion, etc., which is ideal for metal
cutting simulation studies. All these features allow
relatively simple setup of even the most complex
three-dimensional models and make it feasible for
application by industry. Commercial start-up costs
involved are generally in the region of 10 000
15 000, which includes acquisition of the FE software license together with a relatively powerful PC
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

208

S L Soo and D K Aspinwall

or workstation. From a research perspective, however,


these application specific commercial systems have
a number of drawbacks. For example, one of the
major shortcomings of AdvantEdge is that the builtin material property database and solver is a closed
system with no provision for users to implement
custom defined constitutive equations or functions.
This severely restricts the flexibility of the software
and the ability of researchers to troubleshoot models/simulations, which is generally required especially
when investigating new workpiece materials or processes. In a comparative study of three different
commercial FE software (MSC Marc, DEFORM2D &
AdvantEdge) when orthogonally cutting C15 steel,
Bil et al. [95] demonstrated that while satisfactory
agreement was found with certain individual outputs, none of the models were able to show good
correlation with all experimental measured process
parameters. Although, DEFORM2D & MSC Marc
gave reasonable predictions of main cutting force,
AdvantEdge generally over estimated this component, however when considering thrust forces, none
of the simulations were able to provide adequate
results when compared against the experimental
data. The authors attributed the reported discrepancies to the friction models utilized in the respective
software.

ACKNOWLEDGEMENTS
The authors would like to thank James Farrar of
WildeFEA and David OHara of ThirdWave Systems for
providing some of the examples of simulations from
DEFORM3D and AdvantEdge software, respectively.

REFERENCES
1 Tay, A. O., Stevenson, M. G., and de Vahl Davis, G. Using
the finite element method to determine temperature distributions in orthogonal machining. Proc. Instn Mech.
Engrs, 1974, 188(55), 627638.
2 Piispanen,V. Theory of formation of metal chips. J. Appl.
Phys., 1948, 19, 876881.
3 Shaw, M. C. Metal cutting principles, 1984 (Oxford Science Publications, New York).
4 Kalpakjian, S. K. Manufacturing engineering and technology, 3rd edition, 1995 (Addison-Wesley Publishing,
Reading, MA).
5 Merchant, M. E. Mechanics of the metal cutting process
II: plasticity conditions in orthogonal cutting. J. Appl.
Phys., 1945, 16, 318324.
6 Merchant, M. E. Basic mechanics of the metal cutting
process. J. Appl. Mech., 1944, 11, A168A175.
7 Merchant, M. E. Mechanics of metal cutting process 1.
Orthogonal cutting and type 2 chip. J. Appl. Phys., 1945,
16(5), 267275.
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

8 Trent, E. M. and Wright, P. K. Metal cutting, 4th edition,


2000 (Butterworth Heinemann, Boston, MA).
9 Oxley, P. L. B. An analysis for orthogonal cutting with
restricted tool-chip contact. Int. J. Mech. Sci., 1962, 4,
129135.
10 Oxley, P. L. B. A strain-hardening solution for the shear
angle in orthogonal metal cutting. Int. J. Mech. Sci., 1961,
3, 6879.
11 Oxley, P. L. B. and Welsh, M. J. M. Calculating the
shear angle in orthogonal metal cutting from fundamental stressstrainstrain rate properties of the
work material. In Proceedings of the 4th International
Machine Tool Design and Research Conference, 1963,
pp. 7386.
12 Oxley, P. L. B. Mechanics of machining: an analytical
approach to assessing machinability, 1989 (Ellis Horwood
Ltd, Chichester, UK).
13 Trigger, K. J. and Chao, B. T. An analytical evaluation
of metal-cutting temperatures. Trans. ASME, 1951, 73,
5765.
14 Boothroyd, G. Temperatures in orthogonal metal cutting. Proc. Instn Mech. Engrs, 1963, 177(29), 789802.
15 Hou, Z. B. and Komanduri, R. Modelling of thermomechanical shear instability in machining. Int. J. Mech. Sci.,
1997, 39(11), 12731314.
16 Athavale, S. M. and Strenkowski, J. S. Finite
element modeling of machining: from proof-ofconcept to engineering applications. In Proceedings
of the CIRP International Workshop on Modeling
of Machining Operations, Atlanta, USA, May 1998,
pp. 203216.
17 Childs, T. H. C., Maekawa, K., Obikawa, T., and Yamane,
Y. Metal machining, theory and applications, 2000
(Arnold Publishers, London, UK).
18 Movahhedy, M., Gadala, M. S., and Altintas, Y.
Simulation of the orthogonal metal cutting process using an arbitrary LagrangianEulerian finiteelement method. J. Mater. Process. Technol., 2000, 103,
267275.
19 Carroll, J. T. III and Strenkowski, J. S. Finite element
models of orthogonal cutting with application to single point diamond turning. J. Mech. Sci., 1988, 30(12),
899920.
20 Strenkowski, J. S. and Moon, K. J. Finite element prediction of chip geometry and tool-workpiece temperature
distribution in orthogonal metal cutting. Trans. ASME,
J. Eng. Ind., 1990, 112, 313318.
21 Strenkowski, J. S. and Athavale, S. M. A partially constrained Eulerian orthogonal cutting model for chip
control tools. Trans. ASME, J. Manuf. Sci. Eng., 1997, 119,
681688.
22 Kim, K. W. and Sin, H.-C. Development of a
thermo-viscoplastic cutting model using finite element method. Int. J. Mach. Tools Manuf., 1996, 36(3),
379397.
23 Leopold, J. FEM modeling and simulation of 3-D
chip formation. In Proceedings of the CIRP International Workshop on Modeling of Machining Operations,
Atlanta, USA, 1998, pp. 235245.
24 Shet, C. and Deng, X. Finite element analysis of the
orthogonal metal cutting process. J. Mater. Process. Technol., 2000, 105, 95109.
JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

25 Shih, A. J. Finite element analysis of the rake angle effects


in orthogonal metal cutting. Int. J. Mach. Tools Manuf.,
1996, 38(1), 117.
26 Komvopoulos, K. and Erpenbeck, S. A. Finite element
modeling of orthogonal metal cutting. Trans. ASME,
J. Eng. Ind., 1991, 113, 253267.
27 Strenkowski, J. S. and Carroll, J. T. III A finite element
model of orthogonal metal cutting. Trans. ASME, J. Eng.
Ind., 1985, 107, 349354.
28 Ceretti, E. FEM Simulations of segmented chip formation in orthogonal cutting: further improvements.
Proceedings of the CIRP International Workshop on
Modeling of Machining Operations, Atlanta, USA, 1998,
pp. 257263.
29 Ng, E.-G. and Aspinwall, D. K. Modelling of hard
part machining. J. Mater. Process. Technol., 2002, 127,
222229.
30 Marusich, T. D. and Ortiz, M. Modeling and simulation
of high speed machining. Int. J. Numer. Methods Eng.,
1995, 38, 36753694.
31 Wang, J. and Gadala, M. S. Formulation and survey of
ALE method in nonlinear solid mechanics. Finite Elem.
Anal. Des., 1997, 24, 253269.
32 Gadala, M. S., Movahhedy, M. R., and Wang, J.
On the mesh motion for ALE modeling of metal
forming processes. Finite Elem. Anal. Des., 2002, 38,
435459.
33 Movahhedy, M. R., Altintas, Y., and Gadala, M. S.
Numerical analysis of metal cutting with chamfered and
blunt tools. Trans. ASME, J. Manuf. Sci. Eng., 2002, 124,
178188.
34 Joyot, P., Rakotomalala, R., Pantale, O., Touratier, M.,
and Hakem, N. A numerical simulation of steady state
metal cutting. Proc. Instn Mech. Engrs, Part C: J. Mechanical Engineering Science, 1998, 122, 331341.
35 Olovsson, L., Nilsson, L., and Simonsson, K. An
ALE formulation for the solution of two-dimensional
metal cutting problems. Comput. Struct., 1999, 72,
497507.
36 Liang, R. and Khan, A. S. A critical review of experimental
results and constitutive models for BCC and FCC metals
over a wide range of strain rates and temperatures. Int.
J. Plast., 1999, 15, 963980.
37 Lei, S., Shin, Y. C., and Incropera, F. P. Material constitutive modeling under high strain rates and temperatures through orthogonal machining tests. Trans. ASME,
J. Manuf. Sci. Eng., 1999, 121, 577585.
38 Hibbitt, Karlsson & Sorenson, Inc. ABAQUS theory
manual, version 5.8, 1998, USA.
39 Johnson, G. R. and Cook, W. H. A constitutive model
and data for metals subjected to large strains, high
strain rates and high temperatures. In Proceedings of the
7th International Symposium on Ballistics, The Hague,
Netherlands, 1983, pp. 541547.
40 Childs, T. H. C. Material property needs in modelling metal machining. Mach. Sci. Technol., 1998, 2(2),
303316.
41 Soo, S. L., Aspinwall, D. K., and Dewes, R. C. 3D FE
modelling of the cutting of Inconel 718. J. Mater. Process.
Technol., 2004, 150, 116123.
42 Gray, G. T. III Classic split-Hopkinson pressure bar testing. In ASM International metals handbook, mechanical
JMDA163 IMechE 2007

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

209

testing and evaluation, 10th edition, vol. 8, 2000 (ASM


International, Metals Park, OH).
Shi, J. and Liu, C. R. The influence of material models
on finite element simulation of machining. Trans. ASME,
J. Manuf. Sci. Eng., 2004, 126, 849857.
Umbrello, D., MSaoubi, R., and Outeiro, J. C. The
influence of JohnsonCook material constants on finite
element simulation of machining of AISI 316L steel. Int.
J. Mach. Tools Manuf., 2007, 47, 462470.
Xie, J. Q., Bayoumi, A. E., and Zbib, H. M. FEA modeling
and simulation of shear localized chip formation in metal
cutting. Int. J. Mach. Tools Manuf., 1998, 38, 10671087.
Liu, C. R. and Guo, Y. B. Finite element analysis of the
effect of sequential cuts and tool-chip friction on residual
stresses in a machined layer. Int. J. Mech. Sci., 2000, 42,
10691086.
Sasahara, H., Obikawa, T., and Shirakashi, T. FEM
analysis of cutting sequence effect on mechanical characteristics in machined layer. J. Mater. Process. Technol.,
1996, 62, 448453.
Zhang, L. On the separation criteria in the simulation of orthogonal metal cutting using the finite element method. J. Mater. Process. Technol., 1999, 8990,
273278.
Ceretti, E., Fallbohmer, P., Wu, W. T., and Altan, T.
Application of 2D FEM to chip formation in orthogonal
cutting. J. Mater. Process. Technol., 1996, 59, 169180.
Ng, E. G. Modelling of the cutting process when machining hardened steel with polycrystalline cubic boron nitride
(PCBN) tooling. PhD Thesis, School of Manufacturing
and Mechanical Engineering, University of Birmingham,
2001.
Marusich, T. D. and Ortiz, M. Simulation of chip formation in high-speed machining. In Proceedings of the
1995 Joint ASME Applied Mechanics and Materials Conference, Machining of Advanced Materials, AMD, 1995,
vol. 208, pp. 127139.
Obikawa, T., Sasahara, H., Shirakashi, T., and Usui, E.
Application of computational machining method to discontinuous chip formation. Trans. ASME, J. Manuf. Sci.
Eng., 1997, 119, 667674.
Ceretti, E., Lucchi, M., and Altan, T. FEM simulation
of orthogonal cutting: serrated chip formation. J. Mater.
Process. Technol., 1999, 95, 1726.
Hua, J. and Shivpuri, R. Influence of crack mechanics
on the chip segmentation in the machining of Ti-6Al-4V.
In Proceedings of the 9th ISPE International Conference on Concurrent Engineering, Cranfield, UK, 2002,
pp. 357365.
Usui, E. and Takeyama, H. A photoelastic analysis of
machining stresses. Trans. ASME, J. Eng. Ind., 1960, 82,
303308.
Ozel, T. The influence of friction models on finite element simulations of machining. Int. J. Mach. Tools
Manuf., 2006, 46, 518530.
Mamalis, A. G., Horvath, M., Branis, A. S., and
Manolakos, D. E. Finite element simulation of chip formation in orthogonal metal cutting. J. Mater. Process.
Technol., 2001, 110, 1927.
Dirikolu, M. H., Childs, T. H. C., and Maekawa, K. Finite
element simulation of chip flow in metal machining. Int.
J. Mech. Sci., 2001, 43, 26992713.

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

210

S L Soo and D K Aspinwall

59 Li, K., Gao, X.-L., and Sutherland, J. W. Finite element


simulation of the orthogonal metal cutting process for
qualitative understanding of the effects of crater wear on
the chip formation process. J. Mater. Process. Technol.,
2002, 127, 309324.
60 Ko, D. C., Ko, S. L., and Kim, B. M. Rigidthermoviscoplastic finite element simulation of
non-steady-state orthogonal cutting. J. Mater. Process.
Technol., 2002, 130131, 345350.
61 Shi, G., Deng, X., and Shet, C. A finite element study of
the effect of friction in orthogonal metal cutting. Finite
Elem. Anal. Des., 2002, 38, 863883.
62 Lei, S., Shin, Y. C., and Incropera, F. P. Thermomechanical modeling of orthogonal machining process
by finite element analysis. Int. J. Mach. Tools Manuf.,
1999, 39, 731750.
63 Ng, E. G., Aspinwall, D. K., Brazil, D., and Monaghan, J.
Modelling of temperature and forces when orthogonally
machining hardened steel. Int. J. Mach. Tools Manuf.,
1999, 39, 885903.
64 Chandrasekaran, H. and Thuvander, A. Modeling tool
stresses and temperature evaluation in turning using
finite element method. Mach. Sci. Technol., 1998, 2(2),
355367.
65 Monaghan, J. and MacGinley, T. Modelling the orthogonal machining process using coated carbide cutting
tools. Comput. Mater. Sci., 1999, 16, 275284.
66 MacGinley, T. and Monaghan, J. Modelling the orthogonal machining process using coated cemented carbide cutting tools. J. Mater. Process. Technol., 2001, 118,
293300.
67 Guo, Y. B. and Yen, D. W. A FEM study on mechanisms
of discontinuous chip formation in hard machining.
J. Mater. Process. Technol., 2004, 155156, 13501356.
68 Owen, D. R. J. and Vaz, M. Jr Computational techniques
applied to high-speed machining under adiabatic strain
localization conditions. Comput. Methods Appl. Mech.
Eng., 1999, 171, 445461.
69 Bker, M., Rosler, J., and Siemers, C. Finite element simulation of segmented chip formation of Ti6Al4V. Trans.
ASME, J. Manuf. Sci. Eng., 2002, 124, 485488.
70 Bker, M., Rosler, J., and Siemers, C. The influence
of thermal conductivity on segmented chip formation.
Comput. Mater. Sci., 2003, 26, 175182.
71 Maekawa, K., Ohhata, H., Kitagawa, T., and Childs, T.
H. C. Simulation analysis of machinability of leaded CrMo and Mn-B structural steels. J. Mater. Process. Technol.,
1996, 62, 363369.
72 Li, R. and Shih, A. J. Finite element modelling of 3D turning of titanium. Int. J. Adv. Manuf. Technol., 2006, 29,
253261.
73 Guo, Y. B. and Dornfeld, D. A. Finite element modeling of burr formation process in drilling 304 stainless steel. Trans. ASME, J. Manuf. Sci. Eng., 2000, 122,
612619.
74 Bacaria, J.-L., Dalverny, O., and Caperaa, S. A threedimensional transient numerical model of milling. Proc.
Instn Mech. Engrs, Part B: J. Engineering Manufacture,
2001, 215(B8), 11471150.
75 Soo, S. L. 3D modelling when high speed end milling
Inconel 718 superalloy. PhD Thesis, University of
Birmingham, 2003.
Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

76 Miller, S. F. and Shih, A. J. Thermo-mechanical finite


element modeling of the friction drilling process. Trans.
ASME, J. Manuf. Sci. Eng., 2007, 129, 531538.
77 Salio, M., Berruti, T., and De Poli, G. Prediction of residual stress distribution after turning in turbine disks. Int.
J. Mech. Sci., 2006, 48, 976984.
78 Hua, J., Shivpuri, R., Cheng, X., Bedekar, V., Matsumoto,
Y., Hashimoto, F., and Watkins, T. R. Effect of feed rate,
workpiece hardness and cutting edge on subsurface
residual stress in the hard turning of bearing steel using
chamfer + hone cutting edge geometry. Mater. Sci. Eng.,
2005, A394, 238248.
79 Nasr, M. N. A., Ng, E.-G., and Elbestawi, M. A. Modelling
the effects of tool-edge radius on residual stresses when
orthogonal cutting AISI 316L. Int. J. Mach. Tools Manuf.,
2007, 47, 401411.
80 Ee, K. C., Dillon, O. W. Jr, and Jawahir, I. S. Finite element
modeling of residual stresses in machining induced by
cutting using a tool with finite edge radius. Int. J. Mech.
Sci., 2005, 47, 16111628.
81 Zong,W. J., Sun,T., Li, D., Cheng, K., and Liang,Y. C. FEM
optimization of tool geometry based on the machined
near surfaces residual stresses generated in diamond
turning. J. Mater. Process. Technol., 2006, 180, 271278.
82 Chuzhoy, L., Devor, R. E., Kapoor, S. G., and
Bammann, D. J. Microstructure-level modelling of ductile iron machining. Trans. ASME, J. Manuf. Sci. Eng.,
2002, 124(2), 162169.
83 Simoneau, A., Ng, E., and Elbestawi, M. A. Modeling the
effects of microstructure in metal cutting. Int. J. Mach.
Tools Manuf., 2007, 47, 368375.
84 Pramanik, A., Zhang, L. C., and Arsecularatne, J. A. An
FEM investigation into the behaviour of metal matrix
composites: tool-particle interaction during orthogonal
cutting. Int. J. Mach. Tools Manuf., 2007, 47, 14971506.
85 Arola, D. and Ramulu, M. Orthogonal cutting of fiberreinforced composites: a finite element analysis. Int.
J. Mech. Sci., 1997, 39, 597613.
86 Mahdi, M. and Zhang, L. A finite element model for
the orthogonal cutting of fiber-reinforced composite
materials. J. Mater. Process. Technol., 2001, 113, 373377.
87 Arola, D., Sultan, M. B., and Ramulu, M. Finite element
modeling of edge trimming fiber reinforced plastics.
Trans. ASME, J. Manuf. Sci. Eng., 2002, 124, 3241.
88 Venu Gopala Rao, G., Mahajan, P., and Bhatnagar, N.
Micro-mechanical modeling of machining of FRP composites cutting force analysis. Compos. Sci. Technol.,
2007, 67, 579593.
89 Tian, Y. and Shin, Y. C. Multiscale finite element modeling of silicon nitride ceramics undergoing laser-assisted
machining. Trans. ASME, J. Manuf. Sci. Eng., 2007, 129,
287295.
90 Chae, J., Park, S. S., and Freiheit, T. Investigation of
micro-cutting operations. Int. J. Mach. Tools Manuf.,
2006, 46, 313332.
91 Simoneau, A., Ng, E., and Elbestawi, M. A. Chip formation during microscale cutting of a medium carbon steel.
Int. J. Mach. Tools Manuf., 2006, 46, 467481.
92 Liu, X., Devor, R. E., Kapoor, S. G., and Ehmann, K. F.
The mechanics of machining at the microscale: assessment of the current state of the science. Trans. ASME,
J. Manuf. Sci. Eng., 2004, 126(4), 666676.
JMDA163 IMechE 2007

Developments in modelling of metal cutting processes

93 Marafona, J. and Chousal, J. A. G. A finite element model


of EDM based on the Joule effect. Int. J. Mach. Tools
Manuf., 2006, 46, 595602.
94 Bhondwe, K. L., Yadava, V., and Kathiresan, G. Finite
element prediction of material removal rate due to
electro-chemical spark machining. Int. J. Mach. Tools
Manuf., 2006, 46, 16991706.
95 Bil, H., Kilic, S. E., and Tekkaya, A. E. A comparison of
orthogonal cutting data from experiments with three different finite element models. Int. J. Mach. Tools Manuf.,
2004, 44, 933944.

Ft
FN
FS
KIC
p
R
tc
to
V
VC

tangential force in turning (N)


normal force acting on shear
plane (N)
force along shear plane (N)
fracture toughness
hydrostatic stress/pressure (MPa)
resultant force (N)
deformed chip thickness (mm)
undeformed chip thickness (mm)
cutting speed (m/min)
chip velocity (m/min)

APPENDIX

s1

width/thickness of shear zone


strain
strain rate (/s)
temperature ( C)
friction angle ( )
coefficient of friction
normal stress (MPa)
shear stress (MPa)
shear angle ( )

Notation
a
ECSM
EDM
F
Ff
Fr

rake angle ( )
electro-chemical spark machining
electrical discharge machining
frictional force (N)
feed force in turning (N)
radial force in turning (N)

JMDA163 IMechE 2007

211

Proc. IMechE Vol. 221 Part L: J. Materials: Design and Applications

You might also like