Sie sind auf Seite 1von 13

ARTICLE IN PRESS

International Journal of Machine Tools & Manufacture 49 (2009) 586598

Contents lists available at ScienceDirect

International Journal of Machine Tools & Manufacture


journal homepage: www.elsevier.com/locate/ijmactool

Modeling of dynamic micro-milling cutting forces


Mohammad Malekian a, Simon S. Park a,, Martin B.G. Jun b
a

Micro Engineering Dynamics Automation Laboratory (MEDAL), Department of Mechanical and Manufacturing Engineering,
University of Calgary, 2500 University Dr. NW, Calgary, Alberta, Canada T2N 1N4
Laboratory for Advanced Multi-scale Manufacturing, Department of Mechanical Engineering, University of Victoria,
PO Box 3055 STN CSC, Victoria, British Columbia, Canada V8W 3P6
b

a r t i c l e in f o

a b s t r a c t

Article history:
Received 14 May 2008
Received in revised form
10 February 2009
Accepted 14 February 2009
Available online 9 March 2009

This paper investigates the mechanistic modeling of micro-milling forces, with consideration of the
effects of ploughing, elastic recovery, run-out, and dynamics. A ploughing force model that takes the
effect of elastic recovery into account is developed based on the interference volume between the tool
and the workpiece. The elastic recovery is identied with experimental scratch tests using a conical
indenter. The dynamics at the tool tip is indirectly identied by performing receptance coupling analysis
through the mathematical coupling of the experimental dynamics with the analytical dynamics. The
model is validated through micro end milling experiments for a wide range of cutting conditions.
& 2009 Elsevier Ltd. All rights reserved.

Keywords:
Micro-machining
Cutting forces
Receptance coupling
Ploughing

1. Introduction
Highly accurate miniaturized components that are made up
of a variety of engineering materials play key roles in the
future development of a broad spectrum of products [1]. Many
innovative products require higher functionality with signicantly
decreased size; however, conventional fabrication methods using
photolithographic fabrication methods are not applicable to all
engineering materials, and the processes are slow and expensive
and limited to essentially planar geometries [2]. To overcome the
challenges, micro-mechanical machining processes can be utilized
to remove materials mechanically using a miniature tool to create
complex three-dimensional shapes using a variety of engineering
materials [3,4]. Micro-mechanical machining techniques bring
many advantages to the fabrication of micro-sized features. They
can produce micro-components cost-effectively because there is
no need for expensive photolithographic masks. The exibility and
efciency of micro-machining processes using miniature cutting
tools allows for the economical fabrication of smaller batch sizes
compared with other processes [5].
Due to the miniature nature of the mechanical removal
process, micro-machining operations are susceptible to excessive
tool wear, noise, and poor productivity. Thus, the modeling
and understanding of micro-cutting processes are important to
improve the machined part quality and increase productivity.

 Corresponding author. Tel.: +1 4303 220 6959; fax: +1 403 282 8406.

E-mail addresses: mmalekia@ucalgary.ca (M. Malekian),


simon.park@ucalgary.ca (S.S. Park), mbgjun@uvic.ca (M.B.G. Jun).
0890-6955/$ - see front matter & 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmachtools.2009.02.006

However, the conventional mechanistic modeling approach


cannot be applied to micro-scale cutting. In micro end milling,
the cutting edge radius of the end mill is comparable in size to the
chip thickness [6]. As a result, no chip is formed when the chip
thickness is below the minimum chip thickness [7,8]; instead, part
of the work material plastically deforms under the edge of the
tool, and the rest elastically recovers. This change in the chip
formation process, known as the minimum chip thickness effect
and the associated material elastic recovery, causes increased
cutting forces [9] and surface roughness [10] at low feed rates.
Furthermore, when the chip actually forms during cutting with a
nite edge radius tool, ploughing under the edge contributes to an
increase in the specic energy, also known as size effect.
Many researchers have investigated the effect of ploughing on
the size effect. Armarego and Brown [11] suggested that the
greater relative contribution of the ploughing forces with a blunt
tool is responsible for the increase in the specic cutting energy.
Similarly, Lucca et al. [12] showed that the ploughing and elastic
recovery, which were used to explain the increase in the cutting
force, of the workpiece along the ank face of the tool play a
signicant role in micro-machining. Komanduri [13] studied the
ploughing mechanism experimentally by using sharp tools with
extremely negative rake angles to replace the rounded-edge tools.
In order to understand the ploughing mechanisms, ploughing
force models have been developed by many researchers. Vogler
et al. [9] made the rst attempt at incorporating the effect of
minimum chip thickness into a micro end milling force model.
They used the slip-line plasticity model developed by Waldorf
et al. [14]. More complicated slip-line plasticity models that
account for elasticplastic deformation and elastic recovery have

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

Nomenclature
Ap
dx, dy
e
ft
Ft
Fr
Fexp
Ftheo
h
hc
her
Krc, Ktc

ploughed area (mm2)


dynamic tool deection (mm)
error
feed rate (mm/ute)
tangential force (N)
radial force (N)
experimental force (N)
theoretical force (N)
chip thickness (mm)
minimum chip thickness (mm)
height of elastic recovery (mm)
radial and tangential cutting coefcients (N/mm2)

been developed by Jun et al. [15]. Fang [16] also developed a


universal slip-line model for rounded-edge tools. The nite
element model approach has also been utilized by many
researchers to model the micro-cutting process and to understand
size effect [17], machining stresses [18,19], and the inuence of
cutting edge radius on wear resistance [20]. However, the majority
of these methods require many assumptions, and the parameters
used in the model are difcult to estimate. There are a few
mechanistic models developed for micro end milling processes
[2124], but these models do not consider the effects of edge
radius, minimum chip thickness, elastic recovery, and tool
dynamics together.
The objective of this paper is to develop a novel mechanistic
micro-milling cutting force model, based on the shearing and
ploughing-dominant cutting force regimes, that considers different effects, such as elastic recovery, run-out and dynamics.
The mechanistic approach for cutting force modeling has been
very effective for parameter estimation, force prediction, process
monitoring and control, and understanding of the cutting process.
Therefore, development of a new mechanistic micro end milling
force model is important and will be useful for process understanding and monitoring/control.
Micro end mills have small tool tip diameters; therefore,
impact hammer testing cannot be applied directly to the micro
end mills, making it difcult to predict the tool tip dynamics.
To overcome this, the receptance coupling (RC) technique is
employed to mathematically couple the spindle/micro-machine
and arbitrary micro end mills with different geometries in the
prediction of dynamic forces and vibrations.
We rst identify the critical chip thickness, based on the
edge radius and the experimentally obtained forces vs. feed
rate curve. For the shearing-dominant cutting regime, i.e. chip
thickness greater than the critical thickness, we use the conventional sharp-edge theorem to identify the cutting constants by
performing curve ttings from the experimental data. When
the chip thickness is smaller than the critical chip thickness,
we consider a model for the ploughing-dominant cutting regime.
We introduce the ploughing coefcient based on the ploughed
area. The elastic recovery rate of the workpiece is experimentally
identied using a conical indenter and by observing the elastic
deformation after the scratch tests. The cutting force model is
veried for Aluminum 6061 (Al6061).
The organization of the paper is as follows: Section 2
depicts the experimental setup. Section 3 describes the methodology for predicting shearing and ploughing-dominant cutting
forces through mechanistic modeling, including the effects
of the tool tip dynamics, elastic recovery, and run-out and
compares the simulation with the experimentally obtained
forces. Section 4 illustrates the assumptions and limitations

Kre, Kte
Krp, Ktp
N
pe
re
r0
Vp
Xc, Yc
G
H

D
y
ce
ct, cs

587

radial and tangential edge coefcients (N/mm)


radial and tangential ploughing coefcients (N/mm3)
number of utes
elastic recovery (%)
edge radius (mm)
tool run-out (mm)
ploughed volume (mm3)
location of the tool centre (mm)
receptance for the assembled structure (m/N)
receptance for substructures (m/N)
rpm of the spindle (rev/min)
immersion angle (rad)
clearance angle (rad)
geometric angles (rad)

associated with the model, and Section 5 concludes with the


contributions.

2. Experimental setup
We have utilized an ultra precision vertical CNC milling
machine (Kern Micro 2255) with a spindle that can rotate from
60,000 to 160,000 rev/min (rpm). The base of the machine is
polymer concrete, which damps out external vibrations. Unlike
many micro CNC systems, the CNC machine used in this study
utilizes hybrid ball bearings, which provide higher stiffness
and linearity, and an elaborate lubrication system that allows
for temperature stability during the high-speed rotations. The
accuracy of the stage is 1 mm. The experimental setup for this
study is depicted in Fig. 1.
The micro-tools used in this study were uncoated tungsten
carbide (WC) micro end mills with 500 mm diameter at micro
end mills (PMT TS-2-0200-S) with the helix and clearance angles
of approximately 301 and 101, respectively. The tool overhang
length was 15 mm from the collet; and, this value remained
constant so that the dynamics were not changed during the
experiments. The scanning electron microscopy (SEM) picture of
the tip of the 500 mm diameter carbide end mill is shown in Fig. 2.
The edge radius of the tools was measured from the SEM pictures
and observed to be approximately 2 mm.
Several sensors, such as a table dynamometer, an acoustic
emission sensor, accelerometers and capacitance sensors, and an
optical vision system were used to capture various signals and
monitor the cutting processes. The piezo-electric table dynamometer (Kistler 9256C2) with an accuracy of 0.002 N measured
the micro-cutting forces. The charge signals generated from the
force sensor were fed into the charge ampliers (Kistler 9025B),
which converted the charge signals into voltage signals. The
calibration of the table dynamometer was performed using both
modal impact hammer tests (Dytran 5800SL) and a force gauge
(Omega DFG51-2) to verify the force measurement. The sensitivity
of the dynamometer was 26 pC/N for X and Y directions. The noise
level was approximately 0.005 N which was insignicant compared to the cutting forces. The frequency bandwidth of the
dynamometer was found to be approximately 1500 Hz (Fig. 3)
from the impact hammer tests.
The zero point in the Z direction was found by moving
the rotating tool down very slowly and looking at the acoustic
emission (AE) signal carefully. As soon as the tool touched the
workpiece, a sudden jump in the AE signal was observed, and
the position was set to zero. The forces were preprocessed by
subtracting air cutting forces from the measured cutting forces
through synchronization at each revolution of the spindle using

ARTICLE IN PRESS
588

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

Spindle
Capacitance Sensor
Workpiece
MiteeTM Grip

Accelerometer

Interface

AE Sensor

Z
Table Dynamometer
X

XY Stage

Fig. 1. Experimental setup: (a) CNC micro-machine and (b) setup.

Fig. 2. SEM pictures of the tool.

Magnitude [N/N]

15

measure vibration signals in both the X and Y directions. The


workpiece material was Al6061-T6 with a hardness of 95 HB.
The workpiece was attached with the aid of Mitee-GripTM to a
plate that could interface with the dynamometer.

10
3. Cutting force model

Bandwidth

1
0
0

1000

2000

3000 4000
Frequency [Hz]

5000

6000

7000

Fig. 3. Frequency bandwidth of the miniature table dynamometer in the X


direction.

capacitance sensors (Lion Precision C3-D, RD20-2) with an


accuracy of 10 nm. The measured force signals while the spindle
was rotating without material removal (i.e. air cutting) were
subtracted from the cutting forces during the material removal
after synchronizing the two signals based on the capacitance
sensor measurements. An AE sensor (Physical Acoustics Nano30)
was used for capturing high-frequency vibrations. The accelerometers (Kistler 8778A500) were attached to the workpiece to

A mechanistic model has been developed to predict micromilling forces for both shearing and ploughing-dominant cutting
regimes. The model assumes that there is a critical chip thickness
that determines whether a chip will form or not. When machining
is performed at high feed rates, the effects of ploughing and elastic
recovery are insignicant enough to ignore, and the cutting
mechanism is considered to be shearing [10]. However, at lower
feed rates, these effects are substantial and need to be taken into
account. As a result, two different cutting regimes have been
dened. Since the cutting forces in the axial direction are small
compared with the planar directions, only the forces in the X and
Y directions have been considered in this study. The model has
been veried using experimental data for full and half immersion
cutting conditions.
In micro-machining, the edge radius of the tools is considerably large compared to the uncut chip thickness; as a result, the
so-called minimum chip thickness phenomenon occurs in micromachining. Thus, when the uncut chip thickness is less than the
minimum chip thickness (hc), no chip formation occurs and only
ploughing/rubbing takes place. Material separation occurs when

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

h max0; kC ji F ji k  kC ji Ij1
k.
i

shearing cutting coefcients based on the experimental data.


In the ploughing cutting, we introduce ploughing coefcients,
which are the ploughing forces per unit of ploughed volume.
The elastic recovery is identied using the instrumented scratch
tests. The model also considers other effects, such as tool run-out
and the dynamics of the tool at the tool tip that can be obtained
using a capacitance sensor. Since experimental modal analysis
cannot be performed directly on the tip of micro-tools and the

Ploughing Dominant Regime


er

Surface generated
from previous tooth path

hc

Surface generated
from current tooth path
Fig. 5. Chip thickness in the ploughing-dominant regime.

Tooth Path 1

h er
1
Fi

Micro-Endmill

Si+1

Ii
0

Ii+1

i+1

h
D

Ci

Ci+1

The mechanistic micro-milling forces are obtained by considering both the ploughing and shearing effects when the feed rate is
less than the critical chip thickness. We rst determine the

1
Si

Si

from tooth path 0

Fi+1

h
Surface generated

Surface to be generated
from tooth path 1

(1)

Previous Tooth Path

Shearing
Dominant
Regime

Tooth Path 1

the uncut chip thickness is greater than the critical minimum chip
thickness (hc), or at what is sometimes referred to as the
stagnation point [25,26], when the material above the minimum
chip thickness forms a chip and the material below the minimum
chip thickness deforms under the edge with a partial elastic
recovery, resulting in material ploughing. The edge radius of the
tool was 2 mm, and it has been shown that the critical chip
thickness for aluminum is approximately 0.3 of the edge radius
[27,28]. As a result, the critical chip thickness is considered to be
approximately 0.7 mm in this study.
We considered a helix angle of 301 when we discretized the
axial slices in the axial direction. Fig. 4 shows the ploughing
process at the ith rotational angle in micro end milling, when
the chip thickness is less than the minimum chip thickness for an
arbitrary axial slice, where yi represents the angles at the ith
rotational angle, h is the uncut chip thickness, her is the height
of elastic recovery, re is the edge radius, ce is the clearance angle,
and Ap is the ploughed area (represented by the hatched area)
at the rotational angle. The shaded area represents the ploughed
material. The ploughed volume, Vp, at the rotational angle, yi, can
be obtained by summing up the ploughed areas (Ap) of all the
axial slices along the cutting edge.
When the chip thickness, h, increases to greater than the
minimum chip thickness, hc, the ploughing becomes negligible,
and the elastic recovery drops to zero (Fig. 5). Thus, in micro end
milling, each ute goes through different material removal
mechanisms in a single path, and the cutting mechanisms switch
back and forth from the ploughing-dominant regime to the
shearing-dominant regime [28], depending on the uncut chip
thickness value, as shown in Fig. 5.
A comprehensive chip thickness model was developed in [15]
to compute the correct chip thickness, including the effects of the
trochoidal tool path, minimum chip thickness, elastic recovery,
and tool vibrations. Fig. 6 shows the surface generation and chip
thickness computation in the presence of elastic recovery, which
is represented as the shaded region, for an arbitrary axial slice.
Points C and F represent the tool centre and cutting edge locations,
respectively. The superscript denotes the tooth pass number, and
the subscript represents the rotational angle. Point I is found at
the intersection between the previously generated surface from
the previous tooth pass and the line connecting C and F for the
current tooth pass. The chip thickness can be formulated as [15]

589

Surface at the ith rotational angle


Fig. 6. Chip thickness model considering elastic recovery [15].

Tool at i

Ploughed Material

re

her

Current Tooth Path

Ploughed Area Ap

Fig. 4. Ploughing due to the nite edge radius in micro end milling.

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

spindle dynamics cannot be neglected, the receptance coupling


technique is used to nd the tool dynamics.
3.1. Tool dynamics receptance coupling
Micro-tools are very small in diameter; therefore, tool deection can be signicant during the cutting operations. This can
result in excessive tool vibration and cutting forces. In order to
accurately model micro-machining operations, it is important to
predict the deection of the tool subjected to the cutting forces. In
conventional machining, this can be done by accurate measurement of the dynamics, i.e. the frequency response function (FRF)
at the tool tip, using an instrumented hammer and a displacement
sensor. However, impact hammer testing cannot be directly
applied in the determination of the dynamics of the micro end
mills, since the diameter of micro end mills is very small, and the
tools are fragile. Therefore, receptance coupling of the spindle/
machining centre and the micro-tools is employed to extract the
dynamics at the tool tip [29,30]. The RC method allows for the
mathematical coupling of the experimentally and analytically
obtained substructure dynamics to predict the overall assembled
system dynamics. Knowing the coupled FRF at the tool tip,
the deection of the tool due to the cutting forces, and its effects
on the chip formation and surface generation can be further
investigated.
The receptance of a structure is the relationship between
the applied force and the displacement of the structure in the
frequency domain. Knowing the approximate geometry of
the micro end mills, nite element (FE) analysis can be employed
to determine the receptance at the tool tip for different boundary
conditions, such as a cantilever beam condition. However, this
method is not accurate in obtaining the actual FRF of the tool tip,
since the tool is not perfectly clamped and the structural
dynamics of the machine contribute to the FRF of the tool tip
[31]. The receptance coupling method overcomes this challenge
by combining experimentally obtained spindle/machine dynamics
with arbitrary tool dynamics, which are determined with FE
analysis. Fig. 7 shows two substructures: Substructure A consists
of the lower portion of the micro end mill, and Substructure B
includes the remainder of the tool and spindle. The aim of the
receptance coupling, in this case, is to obtain the FRF at the tool tip
(i.e. point 1):
G11

X1
H11  H12 H22 H33 1 H21
F1

where X is a vector with the components of translational (x) and


rotational (y) displacements, and F is a vector with force (f) and
moment (M) components.
The dynamics of Substructure A is obtained using FE analysis,
while the dynamics of Substructure B is determined through
experimental modal analysis. The author [32] has shown that,
for an accurate prediction of the tip dynamics, the rotational
dynamics of the substructures cannot be ignored. Since the
experimental measurement of the rotational dynamics at Location
3 is very difcult, the indirect method as outlined in [29] is used
based on the dynamics measurements using two blank gauge
cylinders (i.e. overhang lengths of 10 and 15 mm from the collet)
of the same material as the tool.
For performing experimental modal analysis to nd the
dynamics of Substructure B, a miniature impact hammer (Dytran
58008L) was used to excite the system. The displacement
was measured using a capacitance sensor (Lion Precision C3D).
The signals were acquired, ltered, and transformed into the
frequency domain using the discrete Fourier transformation. Since
the method is very sensitive to the signal noise, the impact
hammer tests were repeated multiple times and averaged. Also,
the resulting data were curve tted to minimize the effect of
noise.
In this study, the parameters used to formulate the dynamics
of Substructure A with the FE analysis of the tungsten carbide end
mill were: density 14,300 kg/m3, Youngs modulus 580 GPa,
Poissons ratio 0.28, and damping ratio E0.01. The damping
ratio was obtained experimentally by performing the impact
test of blank cylinder with free-free boundary conditions. The
assembled dynamics at the tool tip were obtained using the RC
method in Eq. (2), with the indirectly obtained rotational
dynamics, as depicted in Fig. 8. Two modes were observed in
the frequency range below 10,000 Hz. Therefore, the dynamics of
the system in this range can be written as
G11 jo

o2n;i =ki
o2n;i  o2 2zi on;i jo
i1

2
X

(4)

The modal parameters of the system, i.e. natural frequencies,


stiffness and damping coefcients for the rst and second modes
were obtained through the curve tting method, as shown in
Table 1. The curve tting were performed based on the

x 105

(2)

where G and H denote the assembled and substructure dynamics,


respectively. Since the rotational dynamics cannot be neglected,
the dynamics can be re-represented as
" # "
# "
#
h11;fm
h11;ff
x1
f1
(3)

=
) fX1 g H11 fF1 g
h11;mf h11;mm
y1
M1

Experimental

3
Real [m/N]

590

Curve fitted

2
1
0
1
2
0

1000

2000

3000

5000

6000

5000

6000

4000

x 105

x3

Sub. B

2
x1

M3

Sub. A

M2

f3
f2

Imaginary [m/N]

x2

Experimental

1
0
1
2
3
0

f1

Fig. 7. Receptance coupling of a spindle and an arbitrary end mill [29].

Curve fitted

1000

2000
3000
4000
Frequency [Hz]

Fig. 8. Receptance coupled dynamics (G11) at the tool tip for the 500 mm tool.

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

3.5

Table 1
Dynamical parameters for the tool tip.

k (MN/m)

Second mode

4035
0.016
2.1425

5163
0.038
0.5397

3
2.5

F [N]

minimization of the error between the theoretical and experimentally obtained FRFs using the steepest descent method. The
rst and second modes occur at approximately 4035 and 5163 Hz,
respectively. This dynamic model of the tool allows us to take the
deection of the tool into account during the chip formation
process. If the frequency of the forces are close the natural
frequencies of the system, the forced vibration can be signicant
and immensely affect the micro-cutting operation and surface
generation. It has been assumed that the dynamics of the tool are
equivalent in the X and Y directions, since the tool and spindle
have cylindrical symmetry.
The dynamics of the tool can be utilized to nd the deection
of the tool due to the cutting forces. This deection affects the
surface generation and the cutting forces. If the cutting operation
occurs only in the X direction, the coordinates of the centre of the
end mill can be obtained from:
X Ci f t N

D
60

t i r 0 sin yi dxi ;

Y C i r 0 cos yi dyi

(5)

where ft is the feed rate, N is the number of utes, D is the rpm of


the spindle, y is the rotation angle, r0 is the spindle run-out, and dx
and dy are the dynamic tool deections as a result of cutting forces
in the X and Y directions, respectively. Tool deections that result
from the dynamics are obtained using convolution integrals:
Z t
Z t
F x xg 11 t  xdx; dy
F y xg 11 t  xdx
(6)
dx
0

where Fx and Fy are the cutting forces in the X and Y directions,


respectively, and g11 is the impulse response at the tool tip. Both
modes were considered for the formulation of micro-cutting
forces.
The RC method was utilized to identify the dynamics of the
tool tip for micro end mills, which cannot be acquired directly
using the hammer testing. The method mathematically combines
the dynamics of the micro-tool and the spindle, which were
obtained using FE method and experimental modal analysis,
respectively. The joint rotational dynamics could not be ignored
and were identied indirectly using blank cylindrical tools.
The deection of the tool as a result of tool dynamics, which
can signicantly affect the micro-milling forces and the chip
formation process, is considered in the developed force model.
3.2. Force model development in the shearing-dominant regime
When the chip thickness is bigger than the critical value,
the cutting mechanism is assumed to be similar to the conventional cutting mechanism that considers the shearing and edge
coefcients. In milling operations, the tangential (dFts) and radial
(dFrs) shearing cutting forces acting on a differential ute element
with height dz, as shown in Fig. 4, can be modeled as follows,
when the uncut chip thickness is greater than the minimum chip
thickness value (h4hc) [33]:
dF rs K rc hyi z K re dz
dF ts K tc hyi z K te dz

1.5

Shearing Dominant

Transition

First mode

Ploughing Dominant

on (Hz)

591

1
0.5
0
0

4
6
Feed Rate [m/flute]

10

Fig. 9. Micro-milling resultant forces vs. feed rates at full immersion with depth of
cut of 100 mm.

and, Kre and Kte are the radial and tangential edge coefcients,
respectively. The cutting coefcients represent shearing of the
workpiece, and the edge components represent friction between
the tool and the workpiece. In order to come up with the microcutting force model for the shearing-dominant regime, the
identication of cutting constants is imperative. These coefcients
are obtained from the experimental data; therefore the accurate
measurement of the cutting forces is important. Since the
frequency bandwidth of the table dynamometer is not sufcient
for high-speed cutting operations (i.e. 1500 Hz), the measured
cutting forces are need to be compensated for the unwanted
dynamics of the table dynamometer. A Kalman lter method is
employed as outlined in [34] to accurately measure the highspeed cutting forces based on the dynamics of the sensor as
shown in Fig. 3.
Since the uncut chip thickness is usually bigger than the
critical chip thickness when the feed rate is higher than a critical
value, the cutting mechanism is mainly shearing, and the effect
of ploughing can be neglected. This critical feed rate can
be identied from the forces in this shearing-dominant regime.
Fig. 9 shows the root mean square (RMS) value of the resultant
force vs. feed rate. As can be observed, for feed rates larger than
approximately 2.5 mm/ute, the force has a linear trend. A linear
curve t of the forces in this region is also shown in this gure. In
the ploughing-dominant regime, the forces have increased values
and vary smoothly, but do not follow the same linear behaviour
and are bigger than the predicted values by the extrapolated
linear curve t (dashed line in Fig. 9). A transition region between
the ploughing and shearing regions has been dened, in which
the forces do not follow a smooth curve. The force data in the
shearing-dominant regime can be used to obtain the cutting and
edge coefcients. Since the micro end mills generally have two
utes, the average of the forces is nearly zero, and the ordinary
method [33] of nding the coefcient does not work. Therefore,
the cutting coefcients are found via a nonlinear curve tting. For
this purpose, the following error is minimized through a steepest
descent algorithm [35]:
e

n X
m
X
F expi;j  F theo 2

(8)

i1 j1

(7)

where dz is the height of the differential ute element; Krc and Ktc
are the radial and tangential cutting coefcients, respectively;

where n is the number of feed rates considered in the shearingdominant region, m is the number of samples, Fexp is the
instantaneous experimental forces data (stars in Fig. 9), and Ftheo

ARTICLE IN PRESS
592

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

is the theoretical force obtained from the conventional sharp-edge


theorem using arbitrary initial values for the unknown coefcients. Full immersion cutting data were used for this purpose,
since they showed better agreement with the theoretical data. To
obtain the shearing-dominant cutting coefcients, 8 different feed
rates (n) with 500 samples (m) for each feed rate were used.
Experimental and theoretical forces were synchronized before the
optimization. The coefcients that minimized the error in Eq. (8)
are shown in Table 2.
As can be observed in Table 2, the cutting coefcients were
bigger than those of the conventional end mills, which can be a
result of round edges making the effective rake angle smaller or
even negative, whereas the edge coefcients for the micro end
mills were much less. This is perhaps due to the smaller contact
area between the tool and the workpiece compared with macro
milling operations.
Fig. 10 shows the experimental and theoretical forces obtained
using the identied coefcients for the feed rate of 9 mm/ute and
the full immersion cutting condition in X and Y directions based
on Eq. (7) with the run-out. It can be seen that the conventional
cutting theory with the coefcients shown in Table 2 can predict
the forces in the shearing-dominant regime properly and with
good accuracy. Also, the forces for the feed rate of 9 mm/ute and
the half immersion cutting condition are shown in Fig. 11.
Although the agreement between the experiments and simulations was not as good as what was observed for full immersion,
there was still a good match between the data. The frequency
content of the forces also shows good agreement between the
experiments and the simulation.
The modeling of the cutting forces in micro end milling, similar
to the conventional end milling with the new shearing and edge

Table 2
The cutting constants for the shearing dominant regime.
Parameter

Ktc (N/mm2)

Krc (N/mm2)

Kte (N/mm)

Kre (N/mm)

Al 6061

3650

2420

0.9

0.7

Experiment

coefcients, can appropriately predict the forces in the shearingdominant region. However, it is unable to account for the
increased forces at lower feed rates, and another model for this
ploughing-dominant region needed to be developed.

3.3. Force model development in the ploughing-dominant regime


In the ploughing-dominant regime, the micro-mill undergoes
both ploughing and shearing during the removal process. Chip
formation does not occur when the uncut chip thickness is less
than the minimum chip thickness; instead, there is ploughing and
partial elastic recovery of the material. The ploughing forces are
modeled as proportional to the volume of interference between
the tool and the workpiece. Many researchers have employed this
procedure in the modeling of the forces due to interference
between the clearance face of the tool and the workpiece for
orthogonal cutting [36,37], turning [38,39], and micro end milling
processes [8]. However, none of these modeling approaches have
considered the effect of the elastic recovery during the ploughing
process, as full elastic recovery was assumed for all materials. It
has been shown that the effect of the elastic recovery on the micro
end milling process is substantial [40]. Thus, in this paper, instead
of simply assuming full elastic recovery, the ploughing forces are
modeled as proportional to the volume of interference between
the tool and the workpiece, considering the effect of the elastic
recovery.
Fig. 2 shows two different cases: the height of elastic recovery
(her) is greater than or equal to re(1cos ce), i.e. heXre(1cos ce);
and the height of elastic recovery (her) is less than re(1cos ce),
i.e. herore(1cos ce). The elastic recovery height can be expressed as her peh, where pe is the elastic recovery rate.
The shaded area represents the ploughed area for both cases.
Angle ap is the angle that the point on the rounded edge at the
minimum chip thickness makes with respect to the y-axis, and it
is given by


h
ap cos1 1 
(9)
re

Simulation

Experiment

Simulation

Fy [N]

Fx [N]

1
1
2
3

2
4

4
0

200

400

600 800 1000 1200 1400 1600 1800


Angle [Degree]

400

600

800 1000 1200 1400 1600 1800

Angle [Degree]

12

15

10
8

Magnitude

Magnitude

200

Simulation

6
Experiment

10
Simulation

Experiment

2
0

0
0

2000

4000

6000

Frequency [Hz]

8000

10000

2000

4000

6000

Frequency [Hz]

Fig. 10. Comparison of forces between experiments and simulation at full immersion and a feed rate of 9 mm/ute.

8000

10000

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

2
Experiment

Experiment

Simulation

Simulation

Fy [N]

Fx [N]

1
2

0
1
2

3
4

3
0

200

600

400

800 1000 1200 1400 1600 1800


Angle [Degree]

200

10

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

6
5

Magnitude

Magnitude

593

Simulation

Experiment

4
2

Simulation

Experiment

3
2
1

0
0

2000

4000

6000

8000

10000

2000

Frequency [Hz]

4000

6000

8000

10000

Frequency [Hz]

Fig. 11. Comparison of forces between experiments and simulation at half immersion and a feed rate of 9 mm/ute: (a) X and (b) Y.

Tool at i

Tool at i

re

e
t

re

A
h

C
D

her

her

Fig. 12. Ploughing area for two different cases: (a) herXre(1cos ce) and (b) herore(1cos ce).

For both cases, point B represents the centre of the circle or the
rounded edge, and point D the end of the arc. Line lBD is at the
same angle as the clearance angle from the vertical line. Point A is
at the height of the elastic recovery.
When herXre(1cos ce), the ploughed area, Ap, indicated by the
shaded area in Fig. 12(a), can be obtained as
Ap ABCD AABD  AABC

(10)

where ABCD is the area of the arc segment, and AABD and AABC are
the areas of the triangles connecting the corresponding points.
Area ABCD can be obtained by
ABCD  12r 2e ap ce
The area AABD is given by


r e her  r e 1  cos ce
12r e lAD
AABD
sin ce
2


her  r e 1  cos ce
where lAD
sin ce

(11)

(14)

Thus, the ploughed area Ap can be expressed as


Ap 12r 2e ap ce 12r e lAD  12r e lAB sinap ce ct

(15)

When herore(1cos ce), the ploughed area, Ap, indicated by the


shaded area in Fig. 12(b) can be obtained by
Ap 12r 2e ap cs  sinap cs

(16)

where


cs cos1 1 
(12)

The area AABC can be obtained by


AABC 12r e lAB sinap ce ct

where
q
2
lAB r 2e lAD
 
l
ct tan1 AD
re

(13)

her
re

(17)

For a discretized disk element of a micro end mill ute, the


ploughing volume can be expressed as Vp Apdz, where dz is the
thickness of the disk element. Since the ploughing forces can be
modeled as proportional to the volume of interference between
the tool and the workpiece, both radial and tangential ploughing

ARTICLE IN PRESS
594

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

forces can be written as a multiplication of a constant and the


ploughed volume. Therefore, the ploughing forces for each
differential ute element can be computed as
dF rp K rp Ap K re dz

(18)

dF tp K tp Ap K te dz

where Krp and Ktp are the radial and tangential ploughing coefcients, respectively, and Ap can be determined from Eqs. (15) or
(16). The ploughing coefcients represent ploughing of the
workpiece material. Kre and Kte are the same edge coefcients
as in Eq. (7). Since the edge coefcients represent friction as the
uncut chip thickness goes to zero, the edge components in
the ploughing-dominant regime must be the same as those in the
shearing-dominant regime. The unit for the ploughing coefcients
is N/mm3.

C
b

B
Fig. 13. Finding elastic recovery through an indentation test using a conical
indenter.

Table 3
The cutting constants and estimated parameter values for the ploughing-dominant
regime.
Parameter

Ktp (kN/mm3)

Krp (kN/mm3)

ro (mm)

pe (%)

Al 6061

790

1210

0.2

10

Simulation

0.4

In order to develop a micro-cutting force model for the


ploughing-dominant regime, the identication of ploughing
constants (Krp, Ktp) is essential. The ploughing coefcients in the
ploughing-dominant regime can be identied in the same manner
as the shearing coefcients in the shearing-dominant regime,
based on a nonlinear curve tting with minimized error through a
steepest descent algorithm. In the identication of the ploughing
coefcients, edge coefcients Kre and Kte are set to the values
determined for the shearing-dominant regime. Therefore, the
radial and tangential forces acting on a discretized disk element
can be expressed as
(
dF t
(
dF r

K rc h K re dz

when hXhc shearing

K rp Ap K re dz

when hohc ploughing

0.4

Experiment

(19)

Simulation

Experiment

0.2

Fy [N]

Fx [N]

when hXhc shearing


when hohc ploughing

where h f(ft, y, her, dx, dy, ro), Ap f(h, re, her), ft is the feed per
tooth, and y is the rotational angle of the tool. The computed
forces are summed among all the engaged axial slices over all the
cutting utes to obtain the total tangential and radial forces,
which are then transformed to forces in the planar directions,
with respect to the global coordinate system. The proposed model
suggests cutting and ploughing coefcients that inherently
contain different aspects of plastic deformation, such as strain
hardening and strain gradient effects. The friction forces are
considered constant for different conditions and modeled with
the edge coefcients.
When the uncut chip thickness is smaller than the critical
value, an elastic recovery occurs that can affect the chip
formation, cutting forces, and surface generation during machining operations. The elastic recovery is different for various
materials and should be identied in order to accurately model
the micro-machining operations. It has been shown [41] that the
elastic recovery rate of the material can be identied directly
using instrumented conical scratch tests. In this method, the
remaining grooves from the scratch tests are inspected using a

0.2
0
0.2

0
0.2

0.4

0.4

0.6

0.6

0.8

0.8
0

200

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

1.5

1.5

Magnitude

Magnitude

K tc h K te dz
K tp Ap K te dz

Simulation

Experiment

200

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

Simulation

Experiment

0.5

0.5

0
0

2000

4000
6000
Frequency [Hz]

8000

10000

1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Frequency [Hz]

Fig. 14. Comparison of forces between experiments and simulation at full immersion and a feed rate of 0.5 mm/ute.

ARTICLE IN PRESS

0.4
0.2
0
0.2
0.4
0.6
0.8

Simulation

Simulation

0.4

Experiment

0.2
0
0.2
0.4

200

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

200

400

600 800 1000 1200 1400 1600 1800


Angle [Degree]

2
Magnitude

2.5

Magnitude

595

0.6

Experiment
Fy [N]

Fx [N]

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

2
Simulation

1.5

Experiment

1.5
Simulation Experiment
1
0.5

0.5
0

0
0

2000

6000
4000
Frequency [Hz]

8000

10000

2000

4000

6000

8000

10000

Frequency [Hz]

Fig. 15. Comparison of forces between experiments and simulation at half immersion and a feed rate of 0.5 mm/ute.

0.12

Fx [N]

Magnitude

0.1
0.08

2
Experimental Data
Linear Curve Fit
Simulation

0.06
0

0.04

10

10

0
0

2000

4000
6000
Frequency [Hz]

8000

10000

Fy [N]

0.02
2

Fig. 16. FFT of the AE signal for full immersion and a feed rate of 0.5 mm/ute.

surface prolometer. The surface prole of the groove can be


measured with the aid of the prolometer, as shown in Fig. 13.
The area of the groove under the original surface (abc in Fig. 13)
is then calculated. Knowing this area and the projected area of the
tool in the vertical direction (ABC in Fig. 13), the elastic recovery
rate of the material is acquired from:
pe

SABC  Sabc
SABC

(20)

The scratch tests were performed utilizing conical tools with an


apex angle of 901. The nominal depths of the grooves were chosen
to be 5, 10 and 15 mm; and, the surface prole was measured
at three different points for each groove. It was found that the
average elastic recovery, pe, of Aluminum 6061 was approximately
10%. This acquired elastic recovery was used as one of the
parameters of the developed model for micro-machining.
In order to obtain the ploughing coefcients in Eq. (19), the
same method used for the identication of the coefcients in
the shearing-dominant regime was utilized. The experimental and

6
4
Feed Rate [m/flute]

Fig. 17. Comparison of the RMS of data between experiments and simulation for
full immersion.

theoretical forces were synchronized for full immersion, and


the dened error between the forces (Eq. (8)) was minimized for
10 different feed rates and 500 samples for each feed rate through
the steepest descent method. To assure that the global minimum
was obtained, the initial values of the parameters were varied
several times. As a result, the ploughing coefcients were
identied as shown in Table 3.
Fig. 14 shows the experimental and theoretical forces obtained
using the ploughing-dominant formulation (Eq. (19)) and the
identied coefcients in Tables 2 and 3. It can be observed
that there is good agreement between the simulation and the
experiment; and, the proposed model can properly model the
more complex force proles in the ploughing-dominant
regime. Also, the comparison between the simulations and the

ARTICLE IN PRESS
596

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

Simulation
r0 = 1

0
0.5

0.5
0

200

400

600

Experiment

800

1000 1200 1400 1600 1800

Simulation
r0 = 0.2

Simulation
r0 = 0

200

0.5
Fy [N]

0.5

Fy [N]

Simulation
r0 = 0.2

Experiment

0.5
Fx [N]

Fx [N]

Simulation
r0 = 0

Simulation
r0 = 0.2

Experiment

0.5

0.5

400

600

800

1000 1200 1400 1600 1800


Simulation
r0 = 0.2

Experiment

Simulation
r0 = 1

0
0.5

200

400

600

800

1000 1200 1400 1600 1800

200

400

600

Angle [Degree]

800

1000 1200 1400 1600 1800

Angle [Degree]

Fig. 18. Comparison of forces between experiments and simulation at full immersion and a feed rate of 0.5 mm/ute for different run-outs: (a) r0 0 and 0.2 mm and (b)
r0 0.2 and 1 mm.

Experiment

Simulation
Pe=10%

Simulation
Pe=0%

0.5

Fx [N]

Fx [N]

0.5
0

Simulation
Pe=10%

0
0.5

0.5
0

200

400

600

Experiment

800

1000 1200 1400 1600 1800

Simulation
Pe=10%

0
0.5

200

400

600

Experiment

0.5

Simulation
Pe=0%

Fy [N]

0.5

Fy [N]

Simulation
Pe=20%

Experiment

800

1000 1200 1400 1600 1800

Simulation
Pe=20%

Simulation
Pe=10%

0
0.5

200

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

200

400

600

800 1000 1200 1400 1600 1800


Angle [Degree]

Fig. 19. Comparison of forces between experiments and simulation at full immersion and a feed rate of 0.5 mm/ute for different elastic recoveries: (a) Pe 0% and 10% and
(b) Pe 10% and 20%.

experimental data for half immersion are shown in Fig. 15.


The cutting force predictions in the Y direction show average
deviations of approximately 15% and this can be attributed by
transient vibrations due to the intermittent nature of the half
immersion cutting.
The frequency content of the forces shows that, for low feed
rates, there is more energy at the spindle frequency (1000 Hz),
compared to the forces obtained at higher feed rates because the
effects of run-out and tool imperfections are more signicant
when the feed rate is lower. Also, there is a strong energy
component (especially for forces in the Y direction) near the rst
mode of the system (4000 Hz), which indicates that the second
harmonic of the tooth passing frequency can excite the tool
signicantly and cause forced vibration. This is even more evident

for half immersion conditions, since the second harmonics are


stronger in this situation and cause more excitation at this
frequency. Also, inspection of the frequency component of the AE
sensor (Fig. 16) shows that there is a large frequency component
near the rst mode of the system, which conrms that most of the
vibration happens at the rst natural frequency of the combined
tool/machining centre.
Fig. 17 shows the experimental and theoretical RMS values
of the forces for both the X and Y directions with a linear curve
t of the experimental data for the shearing-dominant region. The
theoretical values were obtained using the developed model with
the identied parameters. It can be observed that the proposed
model can appropriately predict the linear behaviour in the
shearing-dominant regime, as well as the nonlinear behaviour,

ARTICLE IN PRESS
M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

and increased forces in the ploughing-dominant regime. The small


deviation in the data is mainly due to the instabilities that occur
during machining operations.

0.30
0.25

4. Discussions

0.20
alim [mm]

There are several assumptions and limitations associated with


the proposed micro-milling force model. The model in this
paper assumes that the ploughing force is proportional to the
ploughed volume of the material, which is a commonly accepted
assumption by many researchers [8,36]. Also, it is assumed in this
paper that the edge force component is the same, regardless
of the cutting regimes, i.e. shearing and ploughing-dominant
regimes. Thus, the same edge cutting coefcients are used in the
mechanistic models for both the shearing and ploughingdominant regimes. An increase in the friction force due to the
nonlinear interaction between ploughing and rubbing is not
considered in the modeling. In this study, we performed the
experimental tests using uncoated tools, and the frictional forces
may be different for coated tools than for uncoated tools. Also,
the workpiece is assumed to be uniform. Furthermore, tool
deection results in tilted tool conditions that may affect different
parameters, such as rake angle and tool engagement. These
effects, along with thermal effects, have not been considered in
this study.
Utilizing the model developed in this study, the effect of
different parameters, such as elastic recovery and tool run-out,
could be further investigated. Fig. 18 shows the actual and
simulated forces for three different run-outs in full immersion
and a 0.5 mm feed rate. As can be observed in Fig. 18(a), when
there is no tool run-out (r0 0), the maximum of forces is
underestimated but not signicantly, and the agreement between
the forces becomes worse. For a run-out of 1 mm (Fig. 18(b)), there
is a at part in the force simulation, indicating that only one ute
is engaged. The run-out signicantly affects the forces when the
feed rate is small. However, after a certain feed rate, only one ute
is engaged, and the effect of increasing run-out on the cutting
forces is not signicant.
Fig. 19 shows the effect of different elastic recovery for full
immersion cutting at the feed rate of 0.5 mm. It can be observed
that no elastic recovery (0% in Fig. 19(a)) exhibit smaller forces
with higher deviations from the experimental results; while, for
bigger elastic recovery (20% in Fig. 19(b)), slightly higher forces
are predicted. The effects of run-out and elastic recovery in the
shearing-dominant regions are minimal and can be neglected.
The experimental and theoretical forces were also compared
for different depths of cut, which showed good agreement and
veried the validity of the model for different cutting conditions.
The experimental cutting conditions were selected in order to
have chatter-free conditions, by simulating the frequency domain
chatter stability analysis [33] using the mechanistically obtained
cutting parameters. Fig. 20 depicts the chatter stability lobes
of ploughing-dominant full immersion micro-milling operations.
We have selected the chatter-free region at 60,000 rev/min with a
depth of cut of 0.1 mm (shown in Fig. 20). In order to accurately
capture the cutting forces, regenerative chatter-free machining
conditions are imperative [30]. The adverse effect of running the
spindle at the 60,000 rev/min is that forced vibrations may be
experienced.
We have found that the accuracy of the micro-tool fabrications,
with respect to eccentricity, is important to minimize errors [42].
Further studies include investigations of micro-cutting forces
using a single uted cutter and for various workpiece materials,
since different materials behave differently, especially related to
work hardening and size effects.

597

0.15
0.10
0.05
0
0

4
6
Spindle Speed [rev/min]

10
x 104

Fig. 20. Regenerative chatter stability lobes for full immersion micro-milling of
Al6061.

5. Conclusion
The accurate prediction of micro-milling forces is important in
the determination of the optimal machining parameters, in order
to prevent excessive tool wear and poor surface nish while
maintaining high productivity. Unlike macro machining operations, ploughing occurs in micro-machining operations when the
chip thickness is less than the critical chip thickness. There have
been attempts by other researchers to include the effects of
ploughing, elastic recovery and the minimum chip thickness,
based on slip-line plasticity or nite element modeling. However,
these models are very complex, and the estimation of the many
parameters in the models is difcult. In this paper, a mechanistic
force model is developed for the ploughing and shearingdominant regimes for micro end milling operations, considering
run-out, dynamics and the effects of the elastic recovery of
material commonly encountered during micro-machining. Since
the direct measurement of tool tip dynamics is not feasible, we
have employed the receptance coupling method to indirectly
obtain the dynamics. The mechanistic force model has been
veried with experimental cutting force measurements of
Aluminum 6061.

Acknowledgements
This research was supported by the Natural Sciences and
Engineering Research Council of Canada (NSERC), Auto21, and
Canada Foundation for Innovation (CFI). The authors thank Mr.
Abe from Mitsubishi Materials, Japan and Mr. Aarts from Jabro
(Seco) Tools, Netherlands for their generous support.
References
[1] W. Ehrfeld, U. Ehrfeld, Progress and prot through micro technologies.
Commercial applications of MEMS/MOEMS, Proceedings of the SPIE 4561
(2001) 918.
[2] W. Menz, J. Mohr, O. Paul, Microsystem Technology, Wiley-VCH, 2001.
[3] D. Dornfeld, S. Min, T. Takeuchi, Recent advances in mechanical micromachining, Annals of the CIRP 55 (2) (2006) 745768.
[4] Y. Bang, K. Lee, S. Oh, 5-Axis micro milling machine for machining micro parts,
International Journal of Advanced Manufacturing Technology 25 (910)
(2005) 888894.
[5] J. Chae, S.S. Park, T. Freiheit, Investigation of micro-cutting operations,
International Journal of Machine Tools and Manufacture 46 (34) (2006)
313332.

ARTICLE IN PRESS
598

M. Malekian et al. / International Journal of Machine Tools & Manufacture 49 (2009) 586598

[6] G. Bissacco, H.N. Hansen, J. Slunsky, Modelling the cutting edge radius size
effect for force prediction in micro milling, Annals of the CIRP 57 (1) (2008)
113116.
[7] H. Weule, V. Huntrup, H. Tritschle, Micro-cutting of steel to meet new
requirements in miniaturization, Annals of the CIRP 50 (1) (2001) 6164.
[8] X. Lai, H. Li, C. Li, Z. Lin, J. Ni, Modelling and analysis of micro scale milling
considering size effect, micro cutter edge radius and minimum chip thickness,
International Journal of Machine Tools and Manufacture 48 (2008) 114.
[9] M.P. Vogler, S.G. Kapoor, R.E. DeVor, On the modeling and analysis of
machining performance in micro end milling, part II: cutting force prediction,
ASME Journal of Manufacturing Science and Engineering 126 (4) (2004)
695705.
[10] M.P. Vogler, R.E. DeVor, S.G. Kapoor, On the modeling and analysis of
machining performance in micro end milling, part I: surface generation,
ASME Journal of Manufacturing Science and Engineering 126 (4) (2004)
685694.
[11] E.J.A. Armarego, R.H. Brown, On size effect in metal cutting, International
Journal of Production Research 1 (3) (1962) 7599.
[12] D.A. Lucca, R.L. Rhorer, R. Komanduri, Energy dissipation in the ultraprecision
machining of copper, Annals of the CIRP 40 (1) (1991) 6972.
[13] R. Komanduri, Some aspects of machining with negative rake tools simulating
grinding, International Journal of Machine Tool Design 11 (3) (1971) 223233.
[14] D.J. Waldorf, R.E. DeVor, S.G. Kapoor, Slip-line eld for ploughing during
orthogonal cutting, ASME Journal of Manufacturing Science and Engineering
120 (4) (1988) 693698.
[15] M.B.G. Jun, X. Liu, R.E. DeVor, S.G. Kapoor, Investigation of the dynamics of
micro end milling, part 1: model development, ASME Journal of Manufacturing Science and Engineering 128 (4) (2006) 893900.
[16] N. Fang, Slip-line modeling of machining with a rounded-edge tool-part I:
new model and theory, Journal of the Mechanics and Physics of Solids 51 (4)
(2003) 715742.
[17] K. Liu, S.N. Melkote, Material strengthening mechanisms and their contribution to size effect in micro-cutting, ASME Journal of Manufacturing Science
and Engineering 128 (3) (2006) 730738.
[18] K.W. Kim, W.Y. Lee, H.C. Sin, Finite element analysis for the characteristics of
temperature and stress in micro-machining considering the size effect,
International Journal of Machine Tools and Manufacture 39 (9) (1999)
15071524.
[19] T. Ozel, E. Zeren, Finite element modeling of stresses induced by high speed
machining with round edge cutting tools. 2005 ASME International Mechanical Engineering Congress and Exposition, IMECE 2005, November 511, 2005,
Orlando, FL, USA, 2005.
[20] J. Rech, Y.-C. Yen, M.J. Schaff, H. Hamdi, T. Altan, K.D. Bouzakis, Inuence of
cutting edge radius on the wear resistance of Pm-Hss milling inserts, Wear
259 (712) (2005) 11681176.
[21] M.T. Zaman, A.S. Kumar, M. Rahman, S. Sreeram, A three-dimensional
analytical cutting force model for micro end milling operation, International
Journal of Machine Tools and Manufacture 46 (34) (2006) 353366.
[22] W.Y. Bao, I.N. Tansel, Modeling micro end-milling operations. Part I: analytical
cutting force model, International Journal of Machine Tools and Manufacture
40 (15) (2000) 21552173.
[23] I.S. Kang, J.S. Kim, J.H. Kim, M.C. Kang, Y.W. Seo, A mechanistic model of
cutting force in the micro end milling process, Journal of Materials Processing
Technology 187188 (2007) 250255.

[24] H.U. Lee, D.W. Cho, K.F. Ehmann, A mechanistic model of cutting forces in
micro-end-milling with cutting-condition-independent cutting force coefcients, ASME Journal of Manufacturing Science and Engineering 130 (3)
(2008) 03110210311029.
[25] R. Connolly, C. Rubenstein, The mechanics of continuous chip formation in
orthogonal cutting, International Journal of Machine Tool Design and
Research 8 (1968) 159187.
[26] I.S. Jawahir, K.C. Ee, O.W. Dillon Jr., Finite element modeling of residual
stresses in machining induced by cutting using a tool with nite edge
radius, International Journal of Mechanical Sciences 47 (10) (2005)
16111628.
[27] P.K. Basuray, B.K. Misra, G.K. Lal, Transition from ploughing to cutting during
machining with blunt tools, Wear 43 (3) (1977) 341349.
[28] M. Malekian, S.S. Park, Investigation of micro milling forces for aluminum,
Transactions of SME-NAMRI 35 (1) (2007) 417424.
[29] B.A. Mascardelli, S.S. Park, T. Freiheit, Substructure coupling of micro end
mills. ASME International Mechanical Engineering Congress and Exposition,
IMECE2006. Chicago, IL, USA, 2006.
[30] T.L. Schmitz, G.S. Duncan, Receptance coupling for dynamics prediction of
assemblies with coincident neutral axes, Journal of Sound and Vibration 289
(2006) 10451065.
[31] A. Erturk, H.N. Ozguven, E. Budak, Analytical modeling of spindle-tool
dynamics on machine tools using Timoshenko beam model and receptance
coupling for the prediction of tool point FRF, International Journal of Machine
Tools and Manufacture 46 (15) (2006) 19011912.
[32] S.S. Park, Y. Altintas, M. Movahhedy, Receptance coupling for end mills,
International Journal of Machine Tools and Manufacture 43 (9) (2003)
889896.
[33] Y. Altintas, Manufacturing Automation, Cambridge University Press, 2000.
[34] Y. Altintas, S.S. Park, Dynamic compensation of spindle integrated force
sensors, Annals of CIRP 53 (1) (2004) 305309.
[35] P.R. Adby, M.A.H. Dempster, Introduction to Optimization Methods, Chapman
& Hall, London, 1974.
[36] D.W. Wu, A new approach of formulating the transfer function for dynamic
cutting processes, ASME Journal of Engineering for Industry 111 (1989)
3747.
[37] W.J. Endres, R.E. DeVor, S.G. Kapoor, Dual-mechanism approach to the
prediction of machining forces, Part 1: model development, ASME Journal of
Engineering for Industry 117 (4) (1995) 526533.
[38] A.M. Shawky, M.A. Elbestawi, Enhanced dynamic model in turning including
the effect of ploughing forces, ASME Journal of Manufacturing Science and
Engineering 119 (1) (1997) 1020.
[39] M.A. Elbestawi, F. Ismail, R. Du, B.C. Ullagaddi, Modelling machining
dynamics including damping in the toolworkpiece interface, ASME Journal
of Engineering for Industry 116 (4) (1994) 435439.
[40] M.B.G. Jun, R.E. DeVor, S.G. Kapoor, Investigation of the dynamics of micro
end milling, part 2: model validation and interpretation, ASME Journal of
Manufacturing Science and Engineering 128 (4) (2006) 901912.
[41] M. Malekian, S.S. Park, K. Um, Investigation of micro plowing forces
through conical scratch tests, Transactions of SME-NAMRI 36 (1) (2008)
293300.
[42] M.B.G. Jun, K. Bourne, R.E. DeVor, S.G. Kapoor, Estimation of effective error
parameters in high-speed micro-end milling, International Journal of
Machine Tools and Manufacture 47 (9) (2007) 14491454.

Das könnte Ihnen auch gefallen