You are on page 1of 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/238184638

Analysis of Sound Signal Generation Due to Flank Wear in Turning

Article  in  Journal of Manufacturing Science and Engineering · November 2002


DOI: 10.1115/1.1511177

CITATIONS READS
41 124

2 authors, including:

Elijah Kannatey-Asibu
University of Michigan
71 PUBLICATIONS   1,300 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Elijah Kannatey-Asibu on 01 September 2014.

The user has requested enhancement of the downloaded file.


ANALYSIS OF SOUND SIGNAL GENERATION DUE TO FLANK WEAR IN TURNING

Ming-Chyuan Lu
Graduate Student
Elijah Kannatey-Asibu, Jr.
Professor
Department of Mechanical Engineering
The University of Michigan
Ann Arbor, MI 48109-2125

ABSTRACT
Ramp- up is a major step in the implementation of manufacturing systems, and is even more
critical in reconfigurable manufacturing systems. For a successful reduction in ramp-up time, it
is essential to analyze and monitor both the overall manufacturing system and the individual
machine tools/processes that comprise the system. Towards this end, we have addressed the
issue of monitoring tool wear using audible sound to enable faulty conditions associated with
wear to be identified during the process before the part quality gets out of specification. Audible
sound generated from the cutting process is analyzed as a source for monitoring tool wear during
turning, assuming adhesive wear as the predominant wear mechanism. The analysis incorporates
the dynamics of the cutting process. In modeling the interaction on the flank surface, the
asperities on the surfaces are represented as a trapezoidal series function with normal
distribution. The effect of changing asperity height, size, spacing, and the stiffness of the
asperity interaction is investigated and compared with experimental data.

INTRODUCTION
In response to frequent variations in market demand, reconfigurable manufacturing systems
(RMS) are being developed to adapt production systems to capacity and technological changes
[11]. A critical component of RMS is reduction in ramp-up time (the time it takes to get the
production system to produce to full capacity and specification). Ramp-up time reduction is
essential for all manufacturing systems. However, it is even mo re critical with reconfiguration,
since reconfigurable systems are apt to change more often. Thus without the ability to efficiently

1
and quickly identify and resolve errors in the manufacturing system, the benefits of the
reconfigurable system would be defeated. This in turn requires the ability to monitor the overall
system, individual machines, and processes. The stream of variations theory is being developed
to identify root causes of faults associated with the overall manufacturing system [10]. This
involves part measurement to identify where errors are occurring. However, with such an
approach, errors can only be identified after machining is completed at a particular station, and
any damage already done. If the process is monitored, on the other hand, then any potential
problem sources that may be associated with the process, such as tool wear, and which could
affect the part quality can be identified ahead of time, before further damage is done to the part.
Thus process monitoring is a necessary component of ramp- up time reduction methodologies,
and as a consequence, reconfigurable manufacturing systems.
Tool failure monitoring systems that have been developed over the years are
comprehensively reviewed in a number of articles including Micheletti et al. [13], Cook [3],
Tlusty and Andrews [21], Jetly [9], Shiraishi [18], Dan and Mathew [4], Dornfeld [6], and Teti
[20].
Sound monitoring is one technology that has not been extensively investigated for tool
failure monitoring, even though it is extensively used by machine tool operators for decision
making. One major issue related to application of this technology in industry is the ability to
protect the sensor from the hazardous machining environment. However, if appropriate research
can be undertaken to significantly improve the reliability of using sound signals to monitor
machining processes, then production issues can be addressed. Cutting fluids or metal chips
may damage the sensor. This can be prevented by protecting the sensor with a gauze, for
example. Noise from adjacent machines, motors, conveyors, etc. or processes may contaminate
the signals. This effect can be mitigated by using noise cancellation methods in the signal
processing algorithm.
In one of the earliest work on the subject, Weller, Schrier, and Weichbrodt [23] observed that
the total amount of high- frequency vibration energy increases as the length of the cutting edge
wear- land increases. The effects of tool flank wear have also been observed in specific
frequency ranges by Sadat and Raman [17]. Anderson and Dias [2] patented an idea in which
the background noise was sensed by a specific sensor and its effect on signals associated with
tool wear could be reduced. Trabelsi and Kannatey-Asibu [22] obtained audible sound radiation
from a microphone for tool wear and tool breakage detection in turning and determined tool

2
conditions using pattern-recognition techniques. Delio, Tlusty, and Smith [5] concluded that
audible sound is a good method for detecting chatter during the milling process.
In this paper, audible sound is investigated as a basis for tool wear monitoring and a dynamic
model is established to enhance our understanding of the relationship between tool wear and the
sound signal generated from the cutting process. The dynamic system is considered as a forced
vibration system and the interaction between the tool flank and workpiece a source excitation to
the entire system.
Research on the interaction between two sliding surfaces has been reviewed by Ibrahim [7,8],
for friction induced vibration. A smooth spectral density function is used by Soom and Chen
[19] as roughness input model to simulate the random surface roughness- induced contact
vibration at Hertzian contacts during steady sliding. A simple model, which consists of a beam
on an elastic foundation acted upon by a series of moving linear springs, was introduced by
Adams [1] for the dynamic interaction of two dry sliding surfaces. Rice and Moslehy [16]
presented a modeling technique based upon straightforward methods in the study of the
dynamics of mechanical systems, incorporating an interface component of asperities and debris
to simulate dynamic effects in unlubricated sliding.
A lumped model for the dynamics of the cutting process in turning is developed first, and the
system displacement in the cutting and feed directions expressed as functions of the interactive
displacement. The stiffness between the interacting surfaces is assumed to be proportional to
flank wear. Interaction between asperities on the workpiece and flank surface of the cutting tool
is considered here as a source of system excitation. The asperities are modeled as normally
distributed trapezoidal functions.

ANALYSIS

Dynamic System
Since the sound generated during machining is closely tied to the system dynamics, modeling
of the cutting system dynamics was the first step in developing the relationship between tool
wear and the sound signal generated from the cutting process. The system is simplified as a
lumped mass, taking into account the stiffness, damping, cutting force, thrust force, and the force

3
from the interaction between two surfaces of the cutting process. The surface waviness is also
considered in this model. A schematic of the system is shown in Fig. 1.

Workpiece z ?t ? z ?t ? Tt ?

y
z

y x
h c (t)

Workpiece Ff FC

Ft
h c (t): Uncut thickness
C: Damping
Ki K
K: Stiffness Ci M Z

F c : Cutting force

F t : Thrust force Z? Zi CZ
F f : Friction force
z i :Interactive displacement C y Ky
between two contacting surfaces
ssssurfaces
z :Maximum separation during contact
conasperities acontact V
CuttingTool

Fig. 1 Schematic diagram for modeling the cutting process in turning.


The equations of motion of the system shown in Fig. 1 during orthogonal cutting are given
by

.. .
m y ? C y y ? K y y ? Fc ? F f (1a)
.. . .
m z ? C z z ? K z z ? Ft ? C i z i ? K i z i (1b)

where y, z = Displacement of the tool in the y and z directions, respectively

Fc , Ff , Ft = Cutting force, friction force and thrust force, respectively

C y , Cz = Damping coefficients in the y and z directions, respectively

Ci = Damping coefficient associated with contact interaction

K y K z = Stiffness in the y and z directions, respectively

K i = Stiffness associated with the contact interaction

z i = Interactive displacement between two contacting surfaces

m = Mass of the tool.

4
To obtain the system frequency response as a result of tool wear excitation, the interactive
displacement z i , between the two contacting surfaces on the wear land is considered as the input
to the system. Taking the Laplace transform of equations (1) result in equations (2), which
express the governing equations in the s domain. These equations can be solved to obtain the
steady state responses y ?s ?and z ?s ?in terms of z i ?s ?, which reflects the surface characteristics
of the worn tool.

y?s? ?
? ?
hc ?kpc ? ? Dkpz? ms2 ? Czs ? ?Kz ? kpz?e?Tt s ? kpz
s?ms ? C s ? K ?ms ? C s ? K ? k ?e ?1?
2
y y
2
z z pz
?Tt s

(2a)

?
?ms ? C s ? ?k
2
z py
? ?Dkpz ? kpz?e?Tt s ? Kz ? kpy ? kpz??D ?1??Ki ? Ci s? ? ?zi ?s?
?ms ? C s ? K ?ms ? C s ? K ? k ?e
2
y y
2
z z pz
?Tt s
?
?1

z?s ??
? ?
k pz hc ? K i s ? C i s 2 ?z i ?s ?
(2b)
?
s ms ? C z s ? K z ? k pz
2
?e ? Tt s
?1 ??

where
k py , k pz = Dynamic stiffness in the cutting and thrust directions, respectively

? D = Dynamic friction coefficient

Tt =The period between two consecutive cut s at the same locations.

Interactive Displacement z i ?s ?
In the analysis leading to the exciting force function, adhesive wear is considered to be the
dominant mechanism for flank wear, and the asperity interaction as one of the dominant inputs
into the system. Previous work indicates that adhesive wear is always at play whenever one
solid material is slid over the surface of another [14,15]. As tool wear increases and the asperity
tips wear off, the contact area between the flank surface of the tool and the workpiece increases,
while the average asperity height is reduced. The effect of contact action on the cutting system is

5
modeled by considering the asperity interaction between the tool flank and workpiece surface.
Before incorporating the exciting function into the system governing equations, the trapezoid-
shaped asperities were analyzed in the frequency domain. In the analysis, the trapezoidal
function was represented as a general series because the higher frequency content of the signals
is not truncated by using a finite number of terms of the series.
The asperities on the surfaces, represented as the summation of an infinite number of
trapezoids, is shown in Fig. 2, and expressed as z i ?t ? in equation (3):

?
zi ?t ? ? ? A ?U ?t ? At ? ? B ?U ?t ? Bt ? ? C ?U ?t ? Ct ? ? D ?U ?t ? Dt ? (3)
k? 0

where

U ? ? ? Unit step function

A ?
2H
Tk ? 1
? ?
t ? Tk ? ck ??
At ? Tk ? ck

B ? h?
2H
Tk ?1
? ?
t ? T k ? ck ??
h
Bt ? Tk ? ck ? T
2 H k ?1
2H ? ? ? h ? ??
C ? ? ?t ? ?T k ? ck ? ?1 ? ?Tk ? 1 ???
Tk ? 1 ? ?? ? 2 H ? ??

? ? h ? ?
Ct ? ?Tk ? c k ? ?1 ? ?Tk ? 1 ?
? ? 2H ? ?
2H ? ? ? h ? ??
D ? ?h? ?t ? ??Tk ? c k ? ?1 ? ?Tk ? 1 ???
Tk ? 1 ? ? ? 2H ? ??
Dt ? Tk ? c k ? Tk ? 1
T k ? T0 ? T1 ? T2 ? ???Tk ?1
c k ? c 0 ? c1 ? c 2 ? ???c k ?1

c
T

Fig. 2 A series representation of asperities as trapezoids.

6
After taking the Laplace transform, it becomes

? ?
?
zi ?s ? ? ? A?s? ?e? At s ? B?s? ?e? Bt s ? C?s ??e ?Cts ? D?s??e? Dts (4)
k? 0

where
? 2H ? 2H
A?s? ? , B?s?? , C?s? ? , D?s? ?
2H 2H
2 2 2
Tk ?1s Tk ?1s Tk ?1s Tk ?1s2

For simulation, we consider a finite number of terms

? ?
N
zˆi ?s?? ? A?s??e? Ats ? B?s??e? Bt s ? C?s??e?Ct s ? D?s??e? Dt s (5)
k? 0

where
N = Number of terms.

Based on earlier work [12], 100 terms are used in the subsequent simulations.

Forcing Function k i z i ?s ?
In modeling the exciting force on the system, the interaction on the flank surface was
represented by the interactive displacement z i ?t ?, the stiffness k i , and the damping Ci . In the

analysis, k i will be considered to be a variable that depends on tool wear and represented as a
function of the contact area between two contacting surfaces. It is difficult to observe the energy
flow into the system from the analysis of z i ?t ?. By contrast, the stiffness k i is assumed to change
as tool wear increases and represents the mechanism for transferring energy into the system
dynamics. It is represented as a function of the trapezoid top area as follows:

7
2
? ? h ??
k i ? n ?ki ?T ??1 ? ??? (6)
? ? H ??

where

k i = Constant value that depends on the material

n = Constant value associated with the number of contact asperities


T = Average asperity size
h = Average asperity height.

Combining equations (6) and (3) gives

ki ?zi?t?
2
? y? ? (7)
? kˆi ?n???1? m ?? ?? A?U?t? At ?? B?U?t? Bt ??C?U?t?Ct?? D?U?t ?Dt ?
? y ? k?0

where

kˆi ? k i ?T 2

After taking the Laplace transform, we get

k i z i ?s ? ? kˆi ?R?s ?

where

R?s?

? ?
2
? h? ? (8)
? n ???1? ?? ?? A?s??e t ? B?s??e t ? C?s??e t ? D?s??e t
?A s ?B s ?C s ?D s

? H ? k? 0

8
EXPERIMENTAL SETUP
In evaluating the relationship developed, the experimental set up used measured, in addition
to the audible sound signals, other signals such as force, vibration, and acoustic emission. A
Lodge & Shipley turning machine was used to generate different tool wear states. The set-up is
shown in Fig. 3, and features two accelerometers (Montronix VS100 and PCB 480E09), a three
axis dynamometer (Kistler 9257B), and a ½ inch microphone (B&K 4165). The two
accelerometers were installed on the tool holder very close to the tool insert in the feed and
cutting directions, while the microphone was mounted at a distance 7 in away from the cutting
zone, on the end of a fixture supported by a magnetic base on the turret. All signals were
collected in synchrony using a PC and a 12 bit data acquisition board, at a 100 kHz sampling rate
for each channel. An AISI 8620 steel workpiece and an uncoated carbide insert TNMA-432
VC2 were used. Flank wear was measured off- line using a microscope on completing each pass.

Accelerometer PC
Amplifier

Dynamometer Charge
Amplifier

Accelerometer Amplifier Labview

Back Turret
Amplifier
Tool Holder

Microphone
Workpiece
Front
Lathe (Top View)

FIG. 3 Experimental set-up of the multi-sensor system.

RESULTS AND DISCUSSION


In this section, the power spectra of asperities represented as spaced trapezoids are analyzed.
The effect of changing the height y m , spacing c , and size T of the trapezoids on power spectrum
is presented. Each of the three random variables, with a Gaussian distribution, is investigated by

9
keeping the mean of the other two variables constant. The size T is determined by the workpiece
and tool materials, and their respective surface conditions. The amplitude h varies with tool
wear and reduces as wear proceeds, as does the spacing c by considering the fact that the
asperity contact frequency increases as wear proceeds.
From the analysis, we find that a change in the average height h leads to a change in energy
distribution of signals in the frequency domain, but not a change in the dominant signal
frequencies, Fig. 4. As the average asperity height reduces, the energy concentrates on specific
frequencies and the amplitude of the peak is reduced.

x 10
-12
Mean valueh =0.36H -12
Mean valueh =0.66H
x 10
1
1

0.8
0.8
Amplitude

0.6
Amplitude

0.6

0.4
0.4

0.2
0.2

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Frequency (kHz) Frequency (kHz)
(a) (b)

Fig. 4 Variation of surface function spectral characteristics with asperity height.

A change in the average asperity size T leads to a change not only in energy distributions of
signals but also in the location of peaks in the frequency domain. From Fig. 5, it is evident that
as T decreases, the amplitude of the peak is reduced and shifts to a higher frequency.

10
?4
x 10
-13
Mean Value T ? 5.15 ? 10
?4
x 10
-13
Mean Value T ? 8.3 ? 10
1 1

0.8 0.8

Amplitude
Amplitude

0.6 0.6

0.4 0.4

0.2 0.2

0
0
0 1 2 3 4 5 0 1 2 3 4 5

Frequency (kHz) Frequency (kHz)


(a) (b)

Fig. 5 Variation of surface function spectral characteristics with asperity size.

As the spacing c between asperities changes again, the spectral distribution changes not only
in amplitude, but also in the location of the peaks. As c decreases, which corresponds to an
increase in flank wear, the signal shifts to a higher frequency, while also decreasing in peak
amplitude, Fig. 6.
?4
x 10
-13 Mean value c ? 3.95 10 x 10
-13
Mean value c ? 6.82 10
?4
1 1

0.8 0.8
Amplitude

0.6 0.6
Amplitude

0.4 0.4

0.2 0.2

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Frequency (kHz) Frequency (kHz)
(a) (b)

Fig. 6 Variation of spectral distribution with spacing between asperities.

11
When the stiffness of the tool- workpiece interface, k i , is accounted for, the results are

significantly different. As in the previous case, a change in h leads to a change in energy


distribution of signals in the frequency domain, but not a change in the dominant signal
frequencies, Fig. 7. However, as the average asperity height reduces, the amplitude of the peak
frequencies is increased, not decreased as the case in Fig. 4.
Power spectra of signals obtained from the microphone and accelerometers for sharp and
worn tools are shown in Fig. 8. In the frequency domain, the energy distribution for sharp and
worn tools are easily discernible for the sound and vibration signals. Ignoring the sound signals
below 0.5 kHz, similar peaks are observed in both the sound and vibration signals in the feed and
cutting directions, indicating a close relationship between the two types of signals.

Mean valueh ? 0.44H


Mean value h ? 0.56H
-21
x 10 -21
x 10
1.5 1.5

1 1
Amplitude

Amplitude

0.5 0.5

0 0
0 1 2 3 4 5 0 1 2 3 4 5

Frequency (kHz) Frequency (kHz)


(a) (b)
Fig. 7 Variation of spectral characteristics with asperity height when the tool-workpiece
interface stiffness is accounted for.

12
Audible Sound

15
Sharp Tool
Worn Tool

10

Amplitude
5

00 0.5
1 2 3 5
4
Frequency (kHz)

Vibration -Feed Direction Vibration - Cutting Direction


200 700

160 SharpTo ol
SharpTool Worn Tool
WornTool 500
Amplitude

Amplitude

120

80 300

40
100
0 0 0
0 1 2 3 4 5 1 2 3 4 5
Frequency (kHz) Frequency (kHz)

Fig. 8 Audible sound and vibration signals generated during turning with a sharp and a
worn tool. Cutting Speed: 300sfm, Depth of Cut: 0.05in, Feed Rate: 0.005ipr, Flank Wear:
0.017 in.

The power spectral density of the forced vibration system response is expected to include the
characteristics of the exciting input on the flank surface. As discussed above, the power spectral
density of the exciting input, which is represented as a simple trapezoidal series with a normal
distribution, is determined from the analysis to change as tool wear changes. The simulation is

13
verified by comparing it with the sound and vibration signals generated during turning and
observing the trends in the spectral distribution. By comparing the experimental results in Fig. 8
with the simulation results in Figs. 5 and 6, similar trends for changes in the distribution of
power spectral energy are observed, where there is a shift in energy from the lower frequency
range to a higher frequency range as tool wear increases. By including the effect of stiffness of
the interaction between the tool flank and workpiece on energy input, the magnitude of the
system response is also observed to increase as tool wear proceeds. Fig. 9 is the simulation
results obtained when the asperity height, size, and spacing were accounted for simultaneously.

-21
x 10
1.5
Sharp ym ? 5.6?10? 5;
c ? 8.6? 10? 4;T ? 3.4? 10?
1

Worn ym ? 3.6? 10? 5;


c ? 2.6? 10? 4;T ? 3.4?10?
0.5

0
0 1 2 3 4 5
F r e q u en c y (k H z)

Fig. 9 Power spectra of asperities represented as a trapezoidal series function with normal
distribution

CONCLUSIONS
To facilitate ramp-up time reduction, methodologies are being developed to enhance
techniques for monitoring the process to enable faulty conditions to be identified in real time
before undesirable parts are produced. In this regard, sound monitoring is being developed as a
means of enhancing process monitoring capabilities.
In attempting to obtain a better understanding of sound generation as a result of adhesive
wear, asperities on the surfaces of the cutting tool and workpiece are modeled as trapezoids with
normal distribution. The interaction between the asperities is considered to be the source of
system excitation that generates sound signals. It is shown that a change in the asperity form

14
leads to a change in energy distribution of the signal frequencies. Based on the simple model
used, the asperity height is found to decrease as tool wear increases. Under these circumstances,
the dominant and harmonic signals move to higher frequencies. These results agree qualitatively
with the frequency characteristics of experimental data obtained with sharp and worn tools.

ACKNOWLEDGEMENT
The authors are pleased to acknowledge the financial support of the Engineering Research
Center for Reconfigurable Machining Systems (NSF grant #EEC-9529125), and the assistance
from the members of the ERC (both industry and academia).

REFERENCES
1. Adams, G. G., 1996, “Self- Excited Oscillations in Sliding with a Constant Friction
Coefficient- A Simple Model,” ASME Journal of Tribology, Vol. 118, pp. 819-823.
2. Anderson, D., 1988, “Method for Monitoring Cutting Tool Wear During a Machining
Operation,” The Boeing Company, USA, USP. 04744242.
3. Cook, N. H., 1980, “Tool Wear Sensors,” Wear, Vol. 62, pp. 49-57.
4. Dan, L., and Mathew, J., 1990, “Tool Wear and Failure Monitoring Techniques for Turning-
A review,” Int. J. Mach. Tools Manufact., Vol. 30, No. 4, pp. 579-598.
5. Delio, T., Tlusty J., and Smith, S., 1992, “Use of Audio Signals for Chatter Detection and
Control,” ASME Journal of Engineering for Industry, Vol. 114, pp. 146-157.
6. Dornfeld, D. A., 1994, “In Process Recognition of Cutting Stages,” JSME International
Journal, Series C, Vol. 37, No. 4, pp. 638-650.
7. Ibrahim, R. A., 1992a, “Friction-Induced Vibration, Chatter, Squeal, and Chaos: Part 1-
Mechanics of Friction,” Friction-Induced Vibration, Chatter, Squeal, and Chaos, ASME DE-
Vol. 49, pp. 107-121.
8. Ibrahim, R. A., 1992b, “Friction-Induced Vibration, Chatter, Squeal, and Chaos: Part 2-
Dynamic and Modeling,” Friction-Induced Vibration, Chatter, Squeal, and Chaos, ASME,
DE-Vol. 49, pp. 123-138.
9. Jetly, S., 1984, “Measuring Cutting Tool Wear On-Line: Some Practical Considerations,”
Manufacturing Engineering, July, pp. 55-60.

15
10. Jin, J., and Shi, J., 1999, “State Space Modeling of Sheet Metal Assembly for Dimensional
Control,” ASME Journal of Manufacturing Science and Engineering, Vol. 121, pp. 756-762.
11. Koren, Y., Heisel, U., Jovane, F., Moriwaki, T., Pritschow, G., Ulsoy, G., Van Brussel., H.,
1999, “Reconfigurable Manufacturing Systems,” Annals of CIRP, Vol. 48, No. 2, pp. 1-14.
12. Lu, M., and Kannatey-Asibu, Jr., E. , 2000, “Analysis of Sound Signal Characteristics
Associated with Adhesive Wear in Machining,” Transactions of NARMI/SME, Vol. XXVIII,
pp. 257-262.
13. Micheletti, C. F., Koenig, W., and Victor, H. R., 1976, “In Process Tool Wear Sensors for
Cutting Operations,” Annals of CIRP, Vol. 25, pp. 483-496.
14. Rabinowicz, E., 1984, “The Least Wear,” Wear, Vol. 100, pp. 533-541.
15. Rabinowicz, E., 1995, “Adhesive Wear,” Friction and Wear of Materials, John Wiley &
Sons, Inc., New York, pp. 143-190.
16. Rice, S. L., and Moslehy, F. A., 1997, “Modeling Friction and Wear Phenomena,” Wear,
Vol. 206, pp. 136-146.
17. Sadat, A. B., and Raman, S., 1987, “Detection of Tool Flank Wear Using Acoustic Signature
Analysis,” Wear, Vol. 115, pp. 265-272.
18. Shiraishi, M., 1988, “Scope of In-Process Measurement, Monitoring and Control Techniques
in Machining Processes- Part 1: In-Process Techniques for Tools,” Precision Engineering,
Vol. 10, No. 4, pp. 179-188.
19. Soom, A., and Chen, J.W., 1986, “Simulation of Random Surface Roughness-Induced
Contact Vibrations at Hertzian Contacts During Steady Sliding,” Journal of Tribology, Vol.
108, pp. 123-127.
20. Teti, R., 1995, “A Review of Tool Condition Monitoring Literature Data Base,” Annals of
CIRP, Vol. 44/2, pp.659-666.
21. Tlusty, J., and Andrews, G. C., 1983, “A Critical Review of Sensors for Unmanned
Machining,” Annals of CIRP, Vol. 32, No. 2, pp. 563-572.
22. Trabelsi, H., and Kannatey-Asibu, Jr, E., 1991, “Pattern-Recognition Analysis of Sound
Radiation in Metal Cutting,” Int. J. of Advanced Manufacturing Tech., Vol. 6, 220-231.
23. Weller, E. J., Schrier, H. M., and Weichbrodt, B., 1969, “What Sound Can Be Expected from
A Worn Tool,” ASME Journal of Engineering for Industry, Vol. 91, pp. 525-534.

16
APPENDIX A
The equations of motion of the system shown in Fig. 1 during orthogonal cutting are given
by

.. .
m y ? C y y ? K y y ? Fc ? F f (A1a)
.. . .
m z ? C z z ? K z z ? Ft ? C i z i ? K i z i (A1b)

where
y , z = Displacement of the tool in the y and z directions, respectively
Fc , Ff , Ft = Cutting force, friction force and thrust force, respectively

C y , Cz = Damping coefficients in the y and z directions, respectively

Ci = Damping coefficient associated with contact interaction

K y K z = Stiffness in the y and z directions, respectively

K i = Stiffness associated with the contact interaction

z i = Interactive displacement between two contacting surfaces

m = Mass of the tool.

The regenerative effect is considered next with the cutting and thrust forces expressed as
functions of the uncut chip thickness h ?t ? and the friction force in terms of a friction coefficient

? D and the force in the z direction. That gives

Fc ? k py ?z ( t ) ? z ( t ? Tt ) ? h?t ?? (A2a)
F f ? ? D ?Ft ? ?Ki zi ?t ?? Ci z?i ?t ???
(A2b)
? ? D k pz ?z (t ) ? z (t ? Tt ) ? h?t??? ? D ?K i z i ?t?? Ci z?i ?t ??

Ft ? k pz ?z (t ) ? z (t ? Tt ) ? h ?t ?? (A2c)

17
Combining equations (1) and (2), the equations of motion of the system become

m ?y??t ?? Cy y??t?? Ky y?t ? (A3a)


? ?k py ? ? Dk pz ? ?z(t) ? z(t ? Tt ) ? h?t?? ? ? D ?Ki zi ?t ?? Ci ?zi ?t??

m?z??t ? ? Cz z??t ?? K z z ?t ? (A3b)


? k pz ?z(t ) ? z?t ? Tt ? ? h?t ??? Ci z?i ?t ? ? K i z i ?t ?

where
k py , k pz = Dynamic stiffness in the cutting and thrust directions, respectively

? D = Dynamic friction coefficient


Tt = The period between two consecutive cuts at the same location

h = Nominal uncut thickness


z (t ) ? z (t ? Tt ) ? hc ?t ?= Actual uncut thickness.

After taking the Laplace transform of equations (3a) & (3b) we get, in the s domain

?ms ? C s ? K ?y?s?
2
y y

? ?
? ?k pc ? ? Dkpz ?1? e?Tts z ?s ??
?k pc ? ? Dk pz ?h
? ?Ki ? Ci s ?zi ?s ?
(A4a)
s

?ms 2
? C z s ? K z ? k pz ? k pz e
? Tt s
?z ?s ?
k pz h (A4b)
? ? ?K i ? C i s ?z i ?s ?
s

18
By substituting z ?s ? in equation (4b) into equation (4a), y ?s ? and z ?s ? can be expressed as

functions of the interactive displacement z i ?s ? in equation (5).

y?s? ?
?
h?kpc ? ?Dkpz? ms2 ? Czs ? ?Kz ? kpz?e?Tt s ? kpz ??
?
s ms ? Cys ? Ky ms ? Czs ? Kz ? kpz
2
? 2
?e ?1?
?Tts

(A5a)

?
?ms ? C s ? ?k
2
z py ? ?D kpz ? kpz?e?Tt s ? Kz ? kpy ? kpz?? D ?1??Ki ? Cis? ? zi ?s?
?ms ? C s ? K ? ms ? C s ? K ? k ?e
2
y y
2
z z pz
?Tts
?1?

z ?s ? ?
? ?
k pz h ? K i s ? Ci s 2 z i ?s ?
(A5b)
?
s ms ? C z s ? K z
2
? k ?e
pz
? Tt s
?1 ??

which is expressed as equation (2) in the text.

APPENDIX B
The asperities on the surfaces, represented as the summation of an infinite number of
trapezoids, is shown in Fig. 2, and expressed as follows:

?
zi ?t ?? ? A?U?t ? At ?? B?U?t ? Bt ?? C ?U?t ? Ct ?? D?U?t ? Dt ? (B1)
k ?0

where

U ? ?? Unit step function

A?
2H
Tk ? 1
? ?
t ? Tk ? ck ?
At ? Tk ? ck

B ? h?
2H
Tk ? 1
? ?
t ? Tk ? ck ??
h
Bt ? Tk ? ck ? T
2H k?1
2H ? ? ? h ? ??
C?? ?t ? ??Tk ? ck ? ?1 ? ?Tk ? 1 ???
Tk ? 1 ? ? ? 2H ? ??

19
? ? h ? ?
C t ? ?T k ? c k ? ?1 ? ?T k ? 1 ?
? ? 2 H ? ?
2H ? ? ? h ? ??
D ? ?h? ?t ? ?T k ? c k ? ?1 ? ?T k ? 1 ???
T k ? 1 ? ?? ? 2 H ? ??
Dt ? T k ? c k ? T k ?1
T k ? T 0 ? T1 ? T 2 ? ???T k ? 1
c k ? c 0 ? c 1 ? c 2 ? ???c k ? 1

According to the interest in frequency characteristics, the Laplace transform is taken on


equation (B1) that becomes

L?zi ?t ??
?? ?
(B2)
? L?? A?U ?t ? At ? ? B ?U ?t ? Bt ? ? C ?U ?t ? Ct ?? D ?U ?t ? Dt ??
?k ? 0 ?
?
? ? L?A ?U ?t ? A ?? B ?U ?t ? B ? ? C ?U ?t ? C ?? D ?U ?t ? D ??
k?0
t t t t

Each term of equation (B2) can be calculated separately and obtained as follows:

L?A?t ?U ?t ? At ??
? 2H
? L? ?t ? ?Tk ? c k ??U ?t ? ?Tk ? ck ???? (B3a)
?Tk ? 1 ?
2 H ? ?Tk ? c k ?s
? 2 e ? A?s?e ? At s
s T k ?1

L ?B ?t ?U ?t ? Bt ??
?? ???
? L ??? h ?
2H
? ?
t ? T k ? ck ?????U ??? t ? ??Tk ? ck ?
h
T k ? 1 ????
(B3b)
?? Tk ?1 ? ? ? 2H ???
? 2H ? ? h ?? ? ? h ? ??
? L ?? ?? t ? ?Tk ? c k ? T k ? 1 ? ??U ?? t ? ?T k ? c k ? Tk ? 1 ? ???
? Tk ? 1 ? ? 2H ?? ? ? 2H ? ??
? h ?
? 2 H ? ???T k ? c k ? 2 H Tk ? 1 ??? s
? e ? B?s ?e ? Bt s
s 2Tk ? 1

20
L?C?t ?U?t ? Ct ??
?? 2H ? ? (B3c)
??? ? ? ???
?t ? ??Tk ? ck ? ??1? ??Tk ?1????U?t ? ??Tk ? ck ? ??1 ? ??Tk?1????
h h
? L??? ?
???? Tk?1 ? ? ? 2H ? ????? ?? ? ? 2H? ?????
? ? h? ?
? 2H ???Tk ? ck ? ??1? ??Tk ?1 ??s
? e ? ? 2 H? ?
? C?s?e? Ct s
s2Tk? 1

L?D??
t U?t ? Dt ??
?? 2H ? ? h ? ???? ?
? L??? h ?
?
?t ? ??Tk ? ck ? ??1 ?
? ?Tk ?1 ???? U t ? Tk ? ck ? Tk ? 1 ?? ? ??
??? Tk ? 1 ? ? ? 2H ? ???? ?? (B3d)
?2H ?
? L? ? ? ?? ? ?
t ? Tk ? ck ? Tk ? 1 U t ? Tk ? ck ? Tk? 1 ? ??
? k? 1
T ?
2H ? ?Tk ? ck ? Tk ? 1 ?s
? 2 e ? D?s?e ? Dt s

s Tk ? 1

By combining equation (B3d) to (B3d) together, the equation (B2) becomes

? ?
?
zi ?s?? ? A?s? ?e? Ats ? B?s? ?e? Bt s ? C?s??e?Cts ? D?s??e?Dt s (B4)
k ?0

where
? 2H ? 2H
A?s? ? , B?s? ? , C?s? ? , D?s? ?
2H 2H
Tk?1s2 Tk?1s2 Tk?1s2 Tk ?1s2

For simulation, a finite number of terms is considered in equation (B5)

? ?A?s ??e ? (B5)


N
ˆzi ?s ? ? ? At s
? B?s ??e? Bt s ? C ?s? ?e ? Ct s ? D?s??e ? Dts
k? 0

where
N = Number of terms

21
Which is expressed as equation (5) in text.
For futhur analysis, k i is considered as a variable that depends on tool wear to, and can be
represented as a function of the trapezoid top area in equation (B6). The top area is shown in
Fig. B1

a
Atr ? a 2

Fig. B1 Trapezoid top area Atr

k i ? n ?ki a 2
2
? ? h ?? (B6)
? n ?k i ?T ?1 ? ??
? ? H ??

where

k i : Constant value that depends on the material

n : Constant value associated with the number of contact asperities


T : Average asperity size
h : Average asperity height.

Combining equations (B6) and (B1) gives

22
ki ?zi ??
t
? h? ?
2 (B7)
? kˆi ?n???1? ?? ?? A?U?t ? At ? ? B?U?t ? Bt ? ? C?U?t ? Ct ? ? D?U?t ? Dt ?
? H? k ?0

where

kˆi ? k i ?T 2

After taking the Laplace transform, equation(B7) becomes

L?ki zi?t??
? 2
? h? ? ? (B8)
? L?ki ?n ??1? ? ?? A?U?t ? At ? ? B?U?t ? Bt ? ? C?U?t ? Ct ?? D?U?t ? Dt ??
ˆ ? ?
? ? H? k? 0 ?
? ?
2
? h ? ?? ?
? kˆi ?n???1? ?? L?? A?U?t ? At ? ? B?U?t ? Bt ? ? C?U?t ? Ct ?? D?U?t ? Dt ??
? H ? ?k? 0 ?

The L?k i z i ?t ??can represented as:

L?k i z i ?t ??? kˆi ?R ?s ? (B9)

where

? ?
2
? h? ?
R?s?? n ???1? ?? ?? A?s??e? At s ? B?s??e? Bts ? C?s??e? Ct s ? D?s??e?Dt s
? H ? k?0

which is expressed in equation (8) in text.

23
APPENDIX C NOMENCLATURE

Unit Definition

Atr m 2? ? Trapezoid top area


Unit Definition

? ?
C y N ?Sec?m?1 Damping coefficients in the y direction

? ?
Cz N ?Sec?m?1 Damping coefficients in the z direction

Ci ?N ?Sec?m ??1
Damping coefficient associated with contact interaction

c ?m? Spacing between asperities

ck ?m? Summation of spacing through c0 to c k ? 1

c ?m? Mean value of spacing

Fc ?N ? Cutting force

Ff ?N ? Friction force

Ff ?N ? Thrust force

H ?m? Height of asperity associated triangle

h ?m? Average asperity height

hc ?m? Uncut thickness

Ky N ? m? Stiffness in the y direction

Kz ?N m? Stiffness in the y direction

Ki ?N m? Stiffness associated with the contact interaction

Ki ?N m? Contact stiffness associated with the material property

k py ?N m? Dynamic stiffness in the cutting direction

24
k pz N ? m? Dynamic stiffness in the thrust direction

L? ? Laplace transform

m ?Kg ? Mass of tool

n Constant value associated with the number of contact asperities

N Number of terms
Tt ?Sec ? Period between two consecutive cuts at the same locations

T ?m? Asperity size

Tk ?m? Summation of asperity size from T0 to Tk ? 1

T ?m? Average asperity size

U? ? Unit step function


y ?m? Displacement of the tool in the y direction

z ?m? Displacement of the tool in the y direction

zi ?m? Interactive displacement between two contacting surfaces

ẑ i ?m? Interactive displacement between two contacting surfaces with finite term

representation
?D Dynamic friction coefficient

25

View publication stats

You might also like