You are on page 1of 11

60 Int. J. Materials and Product Technology, Vol. 37, Nos.

1/2, 2010

Ultrasonic Assisted Turning of mild steels

Ainhoa Celaya*,
Luis Norberto López de Lacalle,
Francisco Javier Campa
and Aitzol Lamikiz
Faculty of Engineering,
Department of Mechanical Engineering,
University of the Basque Country,
c/Alameda de Urquijo s/n, 48013 Bilbao, Spain
Fax: +34 94 601 4215
E-mail: ainhoa.celaya@ehu.es
E-mail: norberto.lzlacalle@ehu.es
E-mail: fran.campa@ehu.es
E-mail: aitzol.lamikiz@ehu.es
*Corresponding author

Abstract: In this work, the advantages and drawbacks of Ultrasonic Assisted


Turning (UAT) have been investigated focusing on the effect of tool vibration
on surface quality. Several experiments have been performed on mild steels
changing the cutting speed, feed and depth of cut, to study how the influence of
the ultrasonic vibration on the surface roughness varies depending on the
cutting conditions. The results obtained show that the ultrasonic vibration
can improve the surface quality. The authors also propose a new booster design
based on the theory of longitudinal vibration of a bar with varying
cross-sectional area for a higher amplification of the ultrasonic vibration.
Keywords: ultrasonic vibration; turning; roughness; mild steels.

Reference to this paper should be made as follows: Celaya, A.,


López de Lacalle, L.N., Campa, F.J. and Lamikiz, A. (2010) ‘Ultrasonic
Assisted Turning of mild steels’, Int. J. Materials and Product Technology,
Vol. 37, Nos. 1/2, pp.60–70.
Biographical notes: A. Celaya is Lecturer of Manufacturing Processes in the
Department of Mechanical Engineering of the University of the Basque
Country. At present she is carrying out her specialisation studies focused in
Ultrasonic Assisted Turning (UAT) of difficult-to-cut materials. She is also
working in some projects related to High Performance Machining.
L.N. López de Lacalle is Professor of the Department of Mechanical
Engineering of the University of the Basque Country. His research topics
are High Performance Machining, machining of difficult-to-cut materials and
new advances in manufacturing processes.

F.J. Campa is currently working on his PhD in the Department of Mechanical


Engineering of the University of the Basque Country. His research is focused
in the machining of thin-walled parts, dynamics of the milling process and
ultrasonic assisted manufacturing processes.

Copyright © 2010 Inderscience Enterprises Ltd.


Ultrasonic Assisted Turning of mild steels 61

A. Lamikiz is Lecturer of the Department of Mechanical Engineering of the


University of the Basque Country. His main research activity is modelling of
metal removal processes. Laser cutting and Laser polishing are also related
topics working on.

1 Introduction

The Ultrasonic Assisted Turning (UAT) is an assisted machining process in which the
cutting edge of the tool is vibrated at a high frequency around 20 kHz. This technique
has been used both for high precision machining applications and for difficult-to-cut
materials machining such as hardened steels, nickel-based alloys, titanium and
aluminium-SiC metal matrix composites (Zhong and Lin, 2006; Liu et al., 2002).
There are quite a lot of works that have proved the advantages of this technique in terms
of cutting forces, surface finish and tool wear in several materials. Conventional
machining of nickel- and titanium-based alloys presents high tool temperatures and
fast cutting edge wear. High-frequency ultrasonic vibration of the cutting tool has
reportedly allowed a significant reduction of cutting forces and tool wear, a surface finish
improvement up to 25–40%, as well as roundness improvements up to 40–50%
(Babitsky et al., 2003). When cutting low alloy steels (DF2), the ultrasonic vibration
means a reduction in cutting forces up to 50% approximately, and produces
comparatively smaller chips and a better surface finish compared to conventional turning
(Nath et al., 2007). Precision turning of stainless steel has also been tested by applying
ultrasonic vibration to the diamond tool (Moriwaki and Shamoto, 1991). Another field
of application of the technique is the machining of brittle materials. For producing high
quality surfaces on brittle materials, it is important to achieve a ductile deformation
mode. Ultrasonic assisted machining can increase the critical depth of cut at which
ductile-cutting can be obtained. For example, Zhou et al. (2002) machined fused silica
and achieved ductile cutting with a depth of cut up to 1.5 µm where conventional
diamond cutting could only reach 0.5 µm. The tool wear was also reduced by applying
ultrasonic vibration, comparing to conventional turning (Zhou et al., 2006). In short,
there are several advantages of introducing an ultrasonic vibration in the tool:
reduction of cutting forces, shorter chips, surface roughness improvement, tool wear
reduction, roundness improvement, and, as a result, an increase of the machinability of
difficult-to-cut materials.
There are three possible directions in which the ultrasonic vibration can be applied to
the cutting tool in a turning process, which are the cutting speed direction, feed direction
and radial direction. Vibration of the cutting tool in a plane including the cutting speed
direction and the radial direction, known as elliptical vibration, has also been studied.
Elliptical vibration assisted turning is mainly used for precision machining applications
with depths of cut below 50 µm (Moriwaki and Shamoto, 1995; Li and Zhang, 2006).
Nevertheless, it is assumed that the effect of the ultrasonic vibration becomes noticeable
when the vibration velocity is an order of magnitude higher than the velocity of cutting in
the direction of the vibration (Kim and Choi, 1997). So, when the vibration is applied
in the cutting direction, the vibration velocity vt = 2ʌ⋅a⋅fr has to be higher than the cutting
speed v = ʌ⋅n⋅d. And, when the vibration is applied in the feed direction, the vibration
velocity vt = 2ʌ⋅a⋅fr has to be higher than the feed speed f⋅n, where n is the workpiece
62 A. Celaya et al.

rotation speed, d the workpiece diameter, f is the feed, a, the vibration amplitude
and fr the vibration frequency. In this way, during each vibration cycle there is a
separation between the workpiece and the tool, which promotes a pumping action that
decreases the friction (Zhou et al., 2002). Hence, the reduction in the cutting heat and the
work material adherence to the cutting edge of the tool increases the accuracy of the
obtained workpieces (Murakawa and Jin, 1998; Astakhov, 2006).
In several works, when the vibration is applied in the cutting speed direction, the
tested cutting speeds are lower that those applied in the usual industrial applications.
The reason is that, for an ultrasonic vibration with amplitude of 10 µm and a frequency
of 20 kHz, the cutting speed has to be very low for conventional materials, about
75 m/min. However, when the vibration is applied in the feed direction the parameters
used in manufacturing industry can be reached as, usually, feed speeds are below vt.
Consequently, tool vibration in the feed speed direction seems to be more suitable for
industrial ultrasonic turning applications requiring high productivity.
In this work, first, an ultrasonic assisted vibration turning device has been designed
and implemented in a conventional lathe. Then, the device has been tested in order to
investigate the effect of ultrasonic tool vibration on surface quality. Several turning tests
have been done on AISI 1045 steel while changing the cutting speed, the feed and the
depth of cut. Then, the surface roughness obtained with conventional turning and with
UAT has been measured and compared.

2 Experimental procedure

The cutting tests have been performed on an universal lathe. A commercial piezoelectric
transducer Mastersonic MSG-2000 (range of frequencies: 17.5–24 kHz) is used to
generate the ultrasonic vibration. A booster attached to the transducer is used to amplify
the vibration. The design of the booster has to allow the transmission and amplification of
the vibration generated in the transducer while maintaining enough stiffness to be able to
cut the material without a high deflection or dynamic problems. The vibration obtained
has been measured with a laser vibrometer Polytec 0FV-505.
Two configurations have been tested in this work, one applying the ultrasonic
vibration in the cutting speed direction and the other applying ultrasonic vibration in the
feed direction, as shown in Figure 1.

Figure 1 Experimental system when ultrasonic vibration applied in the (a) cutting direction
and (b) feed direction (see online version for colours)

(a) (b)
Ultrasonic Assisted Turning of mild steels 63

For the experiments with vibration in the cutting speed direction, the transducer and the
booster are mounted on a boring bar (S25T-PTFNR11 by Sandvik™) with the proper
overhang. The physic principle is to excite a flexural mode in the range of 17.5–24 kHz.
The measurements with the vibrometer show that a bending vibration mode has been
obtained with maximum vibration amplitude in the tool tip of 8–10 µm at 20.1 kHz.
For the experiments involving the tool vibration in the feed direction, the transducer
has been mounted in the workpiece axis direction, see Figure 1. The transducer excites an
axial mode of the booster, which is clamped at the node of the mode. The measured
resonance frequency of the whole system transducer-booster-tool is 22 kHz and the
amplitude is 10 µm.
A tungsten carbide tool (Sandvik™ TNMG 110304-MF) is employed with a nose
radius of 0.4 mm, 6º rake angle and 6º clearance angle. AISI1045 steel is selected as the
test material. All steel rounds are machined prior to experiments in order to obtain the
same conditions in all the samples and to eliminate any eccentricity.
The same cutting experiments were carried out with and without vibration in order to
compare the cutting performance between the two methods. The tests start with the
application of ultrasonic vibration, and after the tool traverses 20 mm, it continues
without ultrasonic vibration under the same cutting conditions. The following cutting
conditions were used in the experiments: depths of cut from 0.5 mm to 1 mm; feeds from
0.05 mm/rev to 0.3 mm/rev; cutting speeds from 20 m/min to 100 m/min when vibration
is applied in the cutting speed direction and cutting speeds from 164 m/min to 308 m/min
when vibration is in the feed direction. Surface quality was assessed by measuring the
surface average roughness (Ra) and the mean roughness depth (Rz) along the axial
direction of the workpieces. For the obtention of the roughness profiles, a Gaussian filter
with a cut-off length of 0.8 mm and a tracing length of 10 mm was used. Both surfaces
produced by UAT and conventional turning were evaluated using a Taylor Talysurf series
2 surface measuring instrument.

3 Results and discussion

3.1 Surface roughness with vibration in the cutting speed direction


Table 1 shows the effect of ultrasonic vibration on the measured surface roughness,
Ra and Rz. For all cutting conditions, it is generally observed that lower roughness values
are obtained with ultrasonic vibration. In addition, roughness increases with the cutting
speed and depth of cut, as it was expected. In the best of the cases, the average roughness
of the workpiece has improved over 40% when vibration is applied to the cutting tool.
However, the repeatability of the tests was affected by the difficulty to establish
a trustworthy measurement of the amplitude of the vibration during the cutting, and the
loosening of the joints of the system transducer-booster-tool due to the ultrasonic
vibration. Also, the cutting speeds used for the test are slower than those used in the
industry due to the limitation imposed by the vibration velocity. However, as these low
cutting speeds are usual for difficult-to-cut materials, it seems that this alternative has
applications in the machining of those materials.
Figure 2 shows the roughness profiles of the surfaces machined both in conventional
and with vibration cutting. It can be seen that the roughness profile of the UAT test is
64 A. Celaya et al.

more uniform, and for example, it does not reach ±5 µm, as it does the conventional
turning test.

Table 1 Cutting conditions of the tests when vibration is applied in the cutting direction,
and the corresponding surface characteristics

Ultrasonic Cutting speed Depth of cut


vibration Feed (mm/r) (m/min) (mm) Ra (µm) Rz (µm)
Without 0.05 60 0.8 3.530 19.436
With 0.05 60 0.8 1.919 13.780
Without 0.1 41.5 0.5 2.477 13.715
With 0.1 41.5 0.5 1.580 9.760
Without 0.1 50 0.5 2.254 13.35
With 0.1 50 0.5 1.678 9.319
Without 0.1 60 0.5 2.109 12.004
With 0.1 60 0.5 1.719 8.654
Without 0.1 70 0.5 2.301 13.594
With 0.1 70 0.5 1.867 11.264

Figure 2 Roughness component of the machined surface: (a) without ultrasonic vibration
and (b) with ultrasonic vibration (V = 50 m/min; f = 0.1 mm/rev, ap = 0.5) (see online
version for colours)

(a)

(b)
Ultrasonic Assisted Turning of mild steels 65

3.2 Surface roughness with vibration in the feed speed direction


In Table 2, the effect of ultrasonic vibration and turning conditions on the surface
finish of the turned steel samples is shown. In this case, the ultrasonic vibration does not
suppose an appreciable improvement of the roughness, in the best case the improvement
is only 6%. There is also some dispersion in the results and sometimes the use of the
vibration is unfavourable. These poor results can be due to the reduction of the vibration
amplitude during the cutting up to 30–40% depending on the cutting conditions
(Babitsky et al., 2003). The low amplitudes during the cutting must not be enough to
eliminate the crests of the surface generated by the tool. Also the mentioned problems of
loosening in the joints were suffered during these tests.

Table 2 Cutting conditions of the tests when vibration is applied in the feed direction, and the
corresponding surface characteristics

Ultrasonic Cutting speed Depth of cut


vibration Feed (mm/r) (m/min) (mm) Ra (µm) Rz (µm)
Without 0.1 164 0.5 1.107 5.911
With 0.1 164 0.5 1.124 6.657
Without 0.2 164 0.5 2.843 13.94
With 0.2 164 0.5 2.667 13.721
Without 0.2 308 0.5 2.744 12.604
With 0.2 308 0.5 2.659 12.304
Without 0.3 308 0.5 6.559 27.593
With 0.3 308 0.5 6.185 24.624

3.3 Roughness reduction: discussion


For a more in-depth analysis of the surfaces, a spectral analysis of the profiles has been
performed. These kind of techniques can be useful when analysing repetitive or periodic
processes such as turning or milling. For this purpose, the roughness profiles of
Figure 2 have been used as sources for the Fast Fourier Transform (FFT) analysis shown
on Figure 3.
The results shown in Figure 3 indicate differences between the two machined surface
profiles. First, looking at the amplitude of the FFT, the peak corresponding to the rotation
frequency has been reduced with UAT. That means that the crests of the profile left
by the tool feed have been decreased when the vibration was applied to the tool tip.
The analysis also shows a reduction of the amplitude in the higher wavelength
components when ultrasonic vibration is applied, which means that UAT results
in a smoother profile. Finally, at wavelengths shorter than the feed per revolution,
the amplitude of the spectrum is a bit lower in the case of UAT, although it is not easy to
relate that effect directly with the ultrasonic vibration.
But the key question to be discussed is how a high frequency micromovement in the
cutting speed direction can reduce the surface roughness in features characterised by
much lower frequencies. The effect must not be a mechanical effect, because the
micromovement does not affect the roughness direction (is perpendicular to ultrasonic
vibration). Therefore, an indirect way must be proposed.
66 A. Celaya et al.

Figure 3 FFT diagram of the surface profiles of the conventionally machined (CT) and
ultrasonically machined (UAT) workpieces (V = 50 m/min; f = 0.1 mm/rev, ap = 0.5)
(see online version for colours)

In the acoustic monitoring techniques (Lee et al., 2006), the range of frequencies related
to the sliding of metal planes in the primary shear zone is well known. Therefore, it can
be supposed that the micromovement, faster than the cutting speed, varies the mechanism
of plane shearing. At the same time, the effort and pressure of removed chip on the rake
face of tool fluctuates. Therefore, the friction mechanism between the chip and the rake
face also changes, with a reduction of the sticking effect and the secondary shear zone of
the material. The final result could be an easier chip removal process, with a more
defined cutting mechanism (Astakhov, 2006). However, as the roughness is measured in
the feed direction of the turning operation, it is difficult to explain the real mechanism of
roughness reduction. Nevertheless, the improvement on surface quality is present in the
performed tests, and in agreement with other works (Nath et al., 2007; Babitsky et al.,
2004; Wang and Zhao, 1987). On the other hand, when the ultrasonic vibration is applied
along the feed direction, only a slight roughness reduction is observed. Here the basic
hypothesis is simple: at the same time of the feed movement, a very high frequency
movement of the tool edge over the generated crests of workpiece may happen, although
in the experiment it is not observed. The main reason can be that the ultrasonic vibration
movement is not able to penetrate the uncut material, and at the same time the backward
movement does not crush the crests produced by turning.
Anyway, further work must be done in both cases. Also, more effort has to be
dedicated to the optimisation of the transducer-booster-tool system in order to avoid the
loosening of the joints and to fine tune the system to achieve a reliable and repetitive
vibration. That is the way to achieve a roughness reduction that can justify the investment
needed in this assisted process with respect to conventional turning.
Ultrasonic Assisted Turning of mild steels 67

3.4 Redesign of the booster


In order to improve the performance of the system transducer-booster-tool used in the
experimental tests, a new design is proposed. The main objective of the booster is to
increase the amplitude of the vibration generated in the transducer and ensure the
transmission of that vibration to the tool tip. What is more, it is the link between
the transducer-booster-tool system and the lathe, and it has to provide enough stiffness to
the tool in order to avoid deflection problems due to the cutting forces. The booster has to
be designed of such a way that it is resonant with the frequency emitted by the
transducer, so an axial mode has to be found in the range of frequencies provided by
the transducer. The design of the booster has been based on the theory of the longitudinal
vibration of a bar of variable section (Rao, 2004). It is considered an elastic bar of length
L with varying cross-sectional area A(x), as shown in Figure 4. The forces acting on the
cross section of a small element of the bar are given by P and P+dP with
∂u ( x, t )
P = σ ⋅ A( x) = E ⋅ A( x) ⋅ (1)
∂x
where ı is the axial stress, E is the modulus of elasticity, A(x) is the cross-sectional area
at any axial position x, u(x, t) is the longitudinal displacement of the bar section,
and ˜u(x, t)/˜x is the axial strain.

Figure 4 Longitudinal vibration of a bar

The free vibration equation can be obtained by the summation of the forces in the
x direction, where the external forces are zero:
∂ ª ∂u ( x, t ) º ∂ 2 u ( x, t )
« E ⋅ A( x) ⋅ » = ρ ⋅ A( x) ⋅ (2)
∂x ¬ ∂x ¼ ∂t 2
where ȡ is the density of the material.
To solve equation (2), a booster with exponential cross-sectional area is considered:
A( x) = A(0) ⋅ e 2⋅β ⋅ x L (3)

where ȕ is a negative constant.


68 A. Celaya et al.

Supposing a simple harmonic movement of frequency Ȧ, the equation of the


movement is:
u ( x, t ) = ϕ ( x) ⋅ ei ⋅ω ⋅t . (4)

Replacing the expressions of equations (3) and (4) in equation (2), it can be obtained the
following homogenous equation with constant coefficients:
d 2ϕ 2 β dϕ ρω 2
+ + ϕ = 0. (5)
dx 2 L dx E
The solution of equation (5) is:
§ d⋅x d⋅x·
ϕ ( x) = e − β ⋅ x L ¨ A ⋅ cos + B sin ¸ (6)
© L L ¹
where A and B are the integration constants and d = ω 2 L2 ρ E − β 2 .
The previous constants of integration A and B are obtained using the boundary
conditions. Since the transducer of M1 mass is attached to one side of the booster and the
toolholder of mass M2 to the other, the boundary conditions are imposed by the strength
in the bar that must be equal to the force of inertia of the vibrating masses M1 and M2
respectively:
∂u (0, t ) ∂ 2 u (0, t )
A(0) ⋅ E ⋅ = − M 1⋅ (7)
∂x ∂t 2

∂u (0, t ) ∂ 2 u (0, t )
A(0) ⋅ E ⋅ = − M 1⋅ . (8)
∂x ∂t 2
The solution of this differential equation has been programmed in Matlab.
Some parameters of the equations are fixed, for example the properties of the
material, the initial diameter, and the weight of the transducer and the toolholder.
On the other hand, the decay constant of the exponential and the booster length,
have to be selected trough iteration in order to obtain a booster design that amplifies
the vibration amplitude with a frequency inside the range provided by the generator
(17.5–24 kHz). The booster material is AISI 1045 steel, with a modulus of elasticity
of 205GPa and a density of 7870 kg/m3. The dimensions of the initial section are defined
by the transducer section, 50 mm diameter, and the M1 and M2 masses are defined
by the weight of the transducer and toolholder respectively. The obtained exponential
decay constant ȕ value is –0.7 and the booster length is 150 mm. Finally, the frequency
of the first axial mode of the booster is 18.9 kHz and the amplification achieved
is approximately 1 : 3. The resulting geometry and the modal shape are shown in
Figure 5.
The new design of booster has been tested resulting in a surface improvement of 35%
in the best of the cases. For example, applying UAT to medium alloy steel with a feed of
0.1 mm/r, a depth of cut of 0.5 mm and a cutting speed of 70 m/min, the roughness has
been decreased from 2.67 µm Ra to 1.73 µm Ra.
Ultrasonic Assisted Turning of mild steels 69

Figure 5 Final geometry of the designed booster and the corresponding modal shape (see online
version for colours)

4 Conclusions

In the present work, the effect of the ultrasonic vibration in two directions, cutting speed
and feed speed, has been studied. For that purpose, two different vibration devices
have been designed and tested machining mild steel. The experimental tests show an
improvement in surface roughness up to 40% in the most favourable case when the
vibration is applied in the direction of cutting speed and 6% when the vibration is applied
in the feed direction. Also, it has been demonstrated trough an spectral analysis that the
vibration in the cutting speed direction reduces the height of the crests produced by the
tool and reduces the waviness of the generated geometry.
However, there is room for improvement in two aspects. First, the vibration of the
tool tip does not explain completely the reduction of the surface roughness, and a relation
between the vibration and the mechanics of the chip removal has to be established.
The second aspect has a more experimental character, and it is the improvement of the
reliability and the stiffness of the experimental devices. Both problems of joints
loosening and chatter vibrations due to the lack of stiffness were suffered during the
experimentation. Also, more amplitude of the vibration in the tool tip is desirable to reach
a higher vibration velocity. For that purpose, a new booster design, based on the theory of
the longitudinal vibration of a bar of variable section, has been proposed.

Acknowledgement

The authors would like to acknowledge to CIC marGUNE for its collaboration, especially
to Mr. Raul Alberdi, and to the Department of Industry of the Basque Government for its
financial support, as well as to Mr. Eduardo Fraga for his valuable contribution on the
experimental work.
70 A. Celaya et al.

References
Astakhov, V. (2006) Tribology of Metal Cutting, Elsevier, London.
Babitsky, V.I., Kalashnikov, A.N., Meadows, A. and Wijesundara, A.A.H.P. (2003) ‘Ultrasonically
assisted turning of aviation materials’, Journal of Materials Processing Technology, Vol. 132,
pp.157–167.
Babitsky, V.I., Mitrofanov A.V. and Silberschmidt V.V. (2004) ‘Ultrasonically assisted turning
of aviation materials: simulations and experimental study’, Ultrasonics, Vol. 42, pp.81–86.
Kim, J-D. and Choi, I-H. (1997) ‘Micro surface phenomenon of ductile cutting in the ultrasonic
vibration cutting of optical plastics’, Journal of Materials Processing Technology, Vol. 68,
pp.89–98.
Lee, D.E., Hwang, I., Valente, C.M.O., Oliveira, J.F.G. and Dornfeld, D.A. (2006) ‘Precision
manufacturing process monitoring with acoustic emission’, International Journal of Machine
Tools and Manufacture, Vol. 46, pp.176–188.
Li, X. and Zhang, D. (2006) ‘Ultrasonic elliptical vibration transducer driven by single actuator and
its application in precision cutting’, Journal of Materials Processing Technology, Vol. 180,
pp.91–95.
Liu, C.S., Zhao, B., Gao, G.F. and Jiao, F. (2002) ‘Research on the characteristics of the cutting
force in the vibration cutting of a particle-reforced metal matrix composites SiCp/Al’,
Journal of Materials Processing Technology, Vol. 129, pp.196–199.
Moriwaki, T. and Shamoto, E. (1991) ‘Ultraprecision diamong turning of stainless steel by
applying ultrasonic vibration’, Annals of the CIRP, Vol. 40, No. 1, pp.559–562.
Moriwaki, T. and Shamoto, E. (1995) ‘Ultrasonic elliptical vibration cutting’, Annals of the ClRP,
Vol. 44, No. 1, pp.31–34.
Murakawa, M. and Jin, M. (1998) ‘Turning of beta-titanium alloys by means of ultrasonic
vibration’, Trans. NAMRI/SME, Vol. 26, pp.153–158.
Nath, C., Rahman, M. and Andrew, S.S.K. (2007) ‘A study on ultrasonic vibration cutting of low
alloy steel’, Journal of Materials Processing Technology, Vols. 192–193, pp.159–165.
Rao, S.S. (2004) Mechanical Vibrations, Pearson Prentice-Hall, New Jersey, USA.
Wang, L-J. and Zhao J. (1987) ‘Influence on surface roughness in turning with ultrasonic vibration
tool’, Int. J. Mach. Tools Manufact., Vol. 27, No. 2, pp.181–190.
Zhong, Z.W. and Lin, G. (2006) ‘Ultrasonic assisted turning of an aluminium-based metal matrix
composite reinforced with SiC particles’, Int. J. Adv. Manufacturing Technology, Vol. 27,
pp.1077–1081.
Zhou, M., Ngoi, B.K.A., Yusoff, M.N. and Wang, X.J. (2006) ‘Tool wear and surface finish in
diamond cutting of optical glass’, Journal of Materials Processing Technology, Vol. 174,
pp.29–33.
Zhou, M., Wang, X.J., Ngoi, B.K.A. and Gan, J.G.K. (2002) ‘Brittle-ductile transition in the
diamond cutting of glasses with the aid of ultrasonic vibration’, Journal of Materials
Processing Technology, Vol. 121, pp.243–251.

You might also like