You are on page 1of 56

Machining Science and Technology

An International Journal

ISSN: 1091-0344 (Print) 1532-2483 (Online) Journal homepage: https://www.tandfonline.com/loi/lmst20

ULTRASONIC MACHINING—A COMPREHENSIVE


REVIEW

Jatinder Kumar

To cite this article: Jatinder Kumar (2013) ULTRASONIC MACHINING—A


COMPREHENSIVE REVIEW, Machining Science and Technology, 17:3, 325-379, DOI:
10.1080/10910344.2013.806093

To link to this article: https://doi.org/10.1080/10910344.2013.806093

Published online: 22 Jul 2013.

Submit your article to this journal

Article views: 1548

View related articles

Citing articles: 39 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lmst20
Machining Science and Technology, 17:325–379
Copyright # 2013 Taylor & Francis Group, LLC
ISSN: 1091-0344 print=1532-2483 online
DOI: 10.1080/10910344.2013.806093

ULTRASONIC MACHINING—A COMPREHENSIVE REVIEW

Jatinder Kumar
Department of Mechanical Engineering, National Institute of Technology,
Kurukshetra, India

& Ultrasonic machining (USM) is a mechanical material removal process used to erode holes and
cavities in hard or brittle workpieces by using shaped tools, high-frequency mechanical motion, and
an abrasive slurry. The fundamental principles of stationary ultrasonic machining, the material
removal mechanisms involved, proposed models for estimation of machining rate, the effect of
operating parameters on material removal rate, tool wear rate, and workpiece surface finish,
research work reported on rotary mode USM, hybrid USM, process capabilities of USM have been
extensively reviewed in this article. The limitations of USM, gaps observed from the literature
review, and the directions for future research have also been presented. Overall, this article presents
a comprehensive review of USM process for advancement of the process through fundamental
insights into the process.

Keywords future research directions, material removal rate, review, surface quality,
tool wear rate, ultrasonic machining

INTRODUCTION
Ultrasonic machining (USM) is a non-conventional mechanical
material removal process used for machining both electrically conductive
and non-metallic materials; preferably those with low ductility and a
hardness above 40 HRC (Rozenberg, 1973; Saha et al., 1988; Snoyes,
1986; Weller, 1984) such as inorganic glasses, ceramics, quartz, etc. The
process came into existence in 1945 when L. Balamuth was granted the first
patent for the process. USM has been variously termed ultrasonic drilling;
ultrasonic cutting; ultrasonic abrasive machining and slurry drilling.
In USM, high frequency electrical energy is converted into mechanical
vibrations via a transducer=booster combination, which are then trans-
mitted to an energy focusing as well as amplifying device known as horn
or sonotrode. This causes the tool to vibrate along its longitudinal axis at

Address correspondence to Jatinder Kumar, Department of Mechanical Engineering, National


Institute of Technology, Kurukshetra, India. E-mail: jatin.tiet@gmail.com
326 J. Kumar

high frequency; usually greater than 20 kHz with an amplitude of 12–50 mm


(Benedict, 1987; Farago, 1980; Farzin-Nia and Sterrett, 1990; Frederick,
1965; Kennedy and Grieve, 1975; Kremer, 1981; Neppiras, 1964; Pandey
and Shan, 1980). The power ratings range from 50 to 3000 W and a
controlled static load is applied to the tool to provide feed in the
longitudinal direction.
An abrasive slurry, which is a mixture of abrasive material such as silicon
carbide, boron carbide or alumina suspended in water or some suitable car-
rier medium is continuously pumped across the gap between the tool and
work (25–60 mm). The vibration of the tool causes the abrasive particles
held in the slurry to impact the work surface leading to material removal
by micro-chipping (Miller, 1957; Pei and Ferreira, 1995).
Figure 1 shows the basic elements of an USM setup using a magnetos-
trictive or piezoelectric transducer with brazed or screwed tooling. Figure 2
shows the USM setup with an integrated Z stage (Nath et al., 2012).
Variations of this basic configuration include:

1. Rotary Ultrasonic Machining (RUM). In this process, tool vibrates and


rotates simultaneously thereby improving the material removal rate
(MRR) and reducing the geometric inaccuracies; for example, oversize
and out of roundness, etc. (Hu et al., 2002; Li et al., 2005b; Pei and
Ferreira, 1998; Prabhakar and Haselkorn, 1992; Treadwell et al., 2002;

FIGURE 1 Ultrasonic machining head (Weller, 1984).


Ultrasonic Machining 327

FIGURE 2 The USM setup with integrated Z-stage (Nath et al., 2012). (Figure available in color
online.)

Ya et al., 2002; Zeng et al., 2005). The construction of RUM machines is


similar to USM except for the addition of a 0.37–0.56 kW rotary spindle
motor capable of rotating up to 5000 rpm (Perkins, 1972).
2. USM combined with electric discharge machining (Thoe and Aspinwall,
1999; Wansheng et al., 2002; Yan and Biing, 2001; Zhinxin et al., 1995, 2007)
3. Ultrasonic assisted cutting or grinding. Ultrasonic aided turning is the
most common process in this category and is claimed to reduce machin-
ing time, work residual stresses and improve surface quality and tool life
compared to conventional turning (Babitsky et al., 2004; Balamuth,
1966; Chang and Bone, 2005; Ishikawa et al., 1998; Kai and Takahira,
1999; Kohals, 1984; Markov, 1977; Moore, 1985).
4. Other non-machining applications such as cleaning, welding and
coating and polishing.

USM is generally associated with low material removal rates; however its
application is not limited by the electrical or thermal characteristics of the work
material. Because the process is non-thermal and non-chemical, the materials
processed are not altered either chemically or metallurgically (Weller, 1984;
Kumar and Khamba, 2008; Kumar et al., 2012, Kumar et al., 2013a, Kumar
et al., 2013b). For efficient machining to take place, the tool and horn must
be designed with consideration given to mass and shape so that resonance
can be achieved within frequency range capability of the ultrasonic machine.
The purpose of this article is to present a comprehensive review of the
USM process including the construction of ultrasonic machining setup,
328 J. Kumar

tooling considerations, slurry characteristics, process capabilities, mode of


material removal, models proposed by different researchers for prediction
of machining performance, effect of operating parameters on machining
characteristics (such as MRR, tool wear and surface quality), rotary USM,
hybridization of USM with other conventional and non-conventional
machining processes, Lack of understanding in the current research and
directions for future research for advancement of the process. There is
no previously published article that covers all these crucial aspects in a
collective and comprehensive manner for building the fundamental under-
standing of the process. Hence, an attempt has been made to address the
same in this article.

Basic Elements of an Ultrasonic Machine Tool


Ultrasonic machines range from small tabletop-sized units to large
capacity machine tools. In addition to the part-size capacity of a USM
machine, suitability for a particular application is also determined by power
rating (Benedict, 1987). All ultrasonic machines share common subsystems
regardless of the physical size or power (Benedict, 1987). The most impor-
tant of these subsystems are the power supply, transducer element, tool
holder or horn and abrasive supply system. A detailed description of these
components is given as following:

The Ultrasonic Power Supply and Transducer


The power supply for an ultrasonic machine tool is more accurately
characterized as a high power sine wave generator that offers the user
control over both the frequency and power of the generated signal (Benedict,
1987). It converts low frequency (50 Hz) electrical power to high frequency
(approx. 20 kHz). This electrical signal is applied to the transducer for further
conversion into the mechanical motion in form of vibrations. The power sup-
plied depends upon the size of transducer (Balamuth, 1964). Some generators
are designed with safety features such as an auto cut-off in cases of horn frac-
ture, tool failure or overloading (Frederick, 1965).
Transducers convert electrical energy into mechanical motion. With
a conventional generator system, the tool and horn are setup and
mechanically tuned by adjusting their dimensions to achieve resonance.
They can accommodate very small errors in setup or tool wear, giving
minimum acoustic energy loss and very small heat generation
(Kremer, 1981). Two types of transducers used for USM are based on
two different principles of operation, piezoelectric and magnetostrictive
(Benedict, 1987; Kremer, 1981). Because of its lower Q value (Q is a
Ultrasonic Machining 329

measure of the sharpness of the peak value of energy), the


magnetostrictive transducer allows vibration to be transmitted over a wide
frequency band (Kazantsev, 1966). It also allows greater horn design flexi-
bility and can accommodate tool wear. The major drawback of a magnetos-
trictive transducer is high electrical losses and low energy efficiency (about
55–60%). On the other hand, piezoelectric transducers are more energy
efficient (90–96%). Also this type of transducer can generate high vibration
intensities as compared to magnetostrictive transducers (Benedict, 1987;
Farago, 1980; Frederick, 1965; Rozenberg, 1973; Sharma et al., 2003; Xu
and Han, 1999).

The Ultrasonic Horn and Tool Assembly


The function of the tool holder (horn) is to attach and hold the tool
to the transducer. Tool holders are attached to the transducer by means of
a large, loose-fitting screw (Merkulov, 1957). The oscillation amplitude at
the face of the transducer is too small (0.001–0.1 mm) to achieve any
reasonable cutting rate (Amin et al., 2007); therefore, the horn is used
as an amplification device. Half-hard copper washers are used between
the transducer and tool holder to dampen and cushion the interface,
which reduces the chances of any unwanted ultrasonic welding (Rozen-
berg, 1973).
The material used for fabrication of the horn must possess high wear
resistance, good elastic and fatigue strength properties plus corrosion
resistance (Satyanarayna and Krishnan Reddy, 1984). Monel, titanium,
stainless steel and aluminium bronze materials are used for making tool
holders for ultrasonic machines. Figure 3 shows different horn designs with
and without additional tool heads (Thoe et al., 1998). The tool should be
so designed to provide the maximum amplitude of vibration at the free end
at a given frequency (Seah et al., 1993). The material used should have

FIGURE 3 Different horn designs for USM (US 400, 1994).


330 J. Kumar

high wear resistance, good fatigue strength and optimum values of


hardness and toughness (Merkulov, 1957).
Tungsten carbide, silver steel and Monel are the commonly used tool
materials. Polycrystalline diamond (PCD) has recently been detailed for
the machining of very hard work materials such as hot isostatically pressed
silicon nitride (Thoe et al., 1995). Tools can be joined to the horn either by
soldering or brazing, screw fitting. Alternatively, the actual tool configur-
ation can be machined to the end of the horn (Sonic-Mill, 2002). Threaded
joints have also been used conventionally to achieve the ease of tool chan-
ging, however problems such as self-loosening, loss of acoustic power and
fatigue failure have been reported for such tools (Kumehara, 1984).
Depending upon the abrasive used, the work=tool wear ratio can range
from 1:1 to 1:100 (Mishra, 2005). Tool is usually held against the workpiece
by a static load exerted via a counterweight, spring, pneumatic=hydraulic or
solenoid feed system (Kremer, 1981; Pandey and Shan, 1980; Snoyes, 1986).
For optimum results, the system must maintain a uniform working force
while machining and be sufficiently sensitive to overcome the resistance
due to cutting action (Rozenberg, 1973; Rozenberg et al., 1964). Static load
values of about 0.1–30 N are commonly used. This force is particularly criti-
cal when drilling small holes less than 0.5 mm diameter as bending of the
tool can occur under a load that is too high.

Abrasives
In ultrasonic machining, an abrasive slurry (mixture of abrasive and
fluid such as water) is used to achieve the cutting action. Different types
of abrasive materials can be used for making the slurry. Aluminium oxide,
silicon carbide and boron carbide are the most commonly used abrasive
materials. For precision machining and very hard workpiece materials,
cubic boron nitride or diamond powder is also used as the abrasive.
Table 1 details some commonly used abrasive materials and their relative
hardness and cutting abilities.

TABLE 1 Abrasives Used in USM (Mishra, 2005)

Abrasive Knoop hardness Relative cutting power

Diamond 6500–7000 1.0


Cubic Boron Nitride 4700 0.95
Boron carbide (B4C) 2800 0.50–0.60
Silicon carbide (SiC) 2480–2500 0.25–0.45
Alumina (Al2O3) 1850–1920 0.14–0.18

Table 1 and related text in table 3 reproduced from NONCONVENTIONAL


MACHINING, ISBN:81-7319-192-1 # 1997 Narosa Publishing House Pvt Ltd,
New Delhi.
Ultrasonic Machining 331

The transport medium for the abrasive should possess low viscosity with
a density approaching that of the abrasive, good wetting properties and
preferably, high thermal conductivity and specific heat for efficient cooling
(Barash and Watanapongse, 1970; Koops, 1964; Thoe et al., 1998). The most
commonly used concentration is 50% by weight (Kazantsev, 1966). Thinner
mixtures are used to promote efficient flow when drilling deep holes or
forming complex cavities. Once the abrasive is selected and mixed with
water, it is stored in a reservoir at the USM machine and is pumped to the
tool-work interface by the re-circulating pumps at rates up to 36.5 L=min.

Tooling Considerations and Applications


The tool is shaped conversely to the desired hole or cavity and posi-
tioned near, but not touching, the work surface (Benedict, 1987), generally
manufactured by silver brazing (Singh and Khamba, 2003a). In many cases,
a tool of either simple or complex cross-section is penetrated axially to the
workpiece to produce either a through or blind hole of the required
dimensions (Benedict, 1987; Thoe et al., 1998). For three-dimensional
cavities, a process analogous to die sinking is generally employed (Haun
and Schulz, 1994; Moreland, 1987; Neppiras, 1972; Rozenberg et al.,
1964). Figure 4 shows an example of three dimensional cavity forming.
Use of special tools to simultaneously produce a multitude of holes in
precise patterns has also been reported (Moore, 1985). The gang drilling
technique has significantly improved the productivity of USM without com-
promising on quality. Although the volumetric material removal rate in
USM is quite low, the process remains competitive because of its ability with
a single pass of the tool, to generate complex cavities or multiple holes in
work materials that are too hard or fragile to be machined by alternate
processes (Weller, 1984). Application of USM for simultaneous drilling of
multiple holes has been illustrated in Figure 5.

FIGURE 4 Die sinking in silicon nitride turbine blade (Rozenberg and Kazantsev, 1964).
332 J. Kumar

FIGURE 5 USM gang drilling tool for drilling multiple holes (Rutan, 1984).

In a similar application, 2176 square holes measuring 0.79 mm on a side


were simultaneously machined in a 1.01 mm thick carbon plate in 10 min-
utes (Thoe et al., 1998). Because such a large number of holes were drilled
simultaneously, an alternative process that drills one hole at a time would
have to generate holes at rate of 3.6 per second to be able to match the rate
of USM (Benedict, 1987). USM has also been applied for machining of
plastics and engraving of coins. Because of its ability to machine the exact
shape of the tool into the work surface, the accuracy obtained in engraving
operations is exceptional (Gilmore, 1991).
Using ultrasonic machining, a graphite electrode for EDM was shaped
in 30 minutes instead of the 20 hours required by copy milling (Gilmore,
1989; Kremer et al., 1983; Moore, 1985). An alternative approach involves
use of a simple pencil tool in conjunction with a CNC programmed
machine tool to machine the required complex contour in the workpiece.
Figure 6 shows a hypodermic needle that was used to ultrasonically drill
small holes through a silicon nitride workpiece. Recently, this technique
has been investigated in a number of countries including the United
Kingdom, France, Switzerland, Japan etc. (Moreland, 1988a). The
technique removes a large volume of material that would otherwise have

FIGURE 6 Silicon nitride machined by hypodermic needle (Robare and Richerson, 1977).
Ultrasonic Machining 333

been removed by a much slower technique of pantograph profile grinding


(Richardson and Robare, 1978). The selection of USM as a manufacturing
process has occurred because the alternative processes would have required
substantially longer processing times (Benedict, 1987).
Recently, the use of USM for machining of Al2O3=LaPO4 composites
was explored by Majeed et al. (2008). No appreciable surface damage or
defects were reported while machining these composites with USM. In
another investigation (Hocheng and Hsu, 1995), a preliminary study on
ultrasonic drilling of fiber-reinforced plastics was conducted. The machin-
ing efficiency for drilling of multiple holes in FRP has been found to be
superior to other conventional processes, without any structural alterations.

MODELS FOR PREDICTION OF MRR AND MODE OF MATERIAL


REMOVAL
Various analytical models suggested for prediction of material removal
rate thus far (Cook, 1966; Kainth, 1979; Lee and Chan, 1997; Miller, 1957;
Nair and Ghosh, 1985; Rajurkar and Wang, 1999; Rozenberg et al., 1964;
Shaw, 1956; Wiercigroch et al., 1999) have been presented in Table 2. Shaw
(1956) was the first researcher to propose a static and simple analytical
model giving the relationship between MRR and vibration amplitude, fre-
quency, abrasive grit size and concentration and feed force. However, its
predictions do not agree well with the experimental observations. Miller
(1957) proposed a MRR model based on the plastic deformation restricting
its application to ductile materials only, while Rosenberg et al. (1964)
included the statistical distribution of abrasive particle size in their compu-
tational model.
Cook (1966) developed the simplest model to predict the linear
machining rate. Kainth et al. (1979) proposed another static, complicated
model using the abrasive particle size distribution. Nair and Ghosh (1985)
also proposed a computationally intensive model simulating the princi-
ples of elastic wave propagation. Wang and Rajurkar (1999) suggested a
more realistic model, taking into account the stochastic and dynamic nat-
ure of the process. But this is applicable to perfectly brittle materials only.
Lee and Chan (1997) developed an analytical model to predict the effects
of vibration amplitude, grit size and feed force on MRR and surface
roughness for ceramic composites. Wiercigroch et al. (1999) have pro-
posed another model to predict the MRR in ultrasonic drilling using
an impact oscillator approach. The inherent non-linearity of the discon-
tinuous impact process has been modelled to generate the pattern of
impact forces. This model explains the experimentally observed fall in
MRR at higher static forces.
334 J. Kumar

TABLE 2 MRR Models for USM (Jain and Jain, 2001)


Name of Mechanism of material
Investigator removal Assumptions Limitations

M.C. Shaw Direct hammering of All abrasive particles are Analysis does not agree
(1956) abrasive particle identical, rigid, spherical in with experimental
(Primary) shape. results qualitatively.
Impacting by free moving All impacts are identical. Does not predict the effect
particles (secondary) Material removal is proportional of variation in
to volume of material amplitude, feed force or
removed per impact, number frequency correctly.
of particles impacting per Predicts infinite increase
cycle and frequency of in machining rate with
impacting. static force while an
Penetration depth is inversely optimum value exists
proportional to flow stress of due to grain crushing.
work material. No allowance for grain size
For a given area of tool face, variation and for
number of active grains is crushed grains.
inversely proportional to
square of the mean diameter
of grains.
G.E. Miller By chipping plastically Abrasive particles are of cubical Applicable to ductile
(1957) deformed and work size. materials only as MRR is
hardened material. In Plastic deformation is directly assumed to depend on
ductile material, MRR proportional to the stress. plastic deformation.
depends upon work Plastic flow stress equals Burger Some non-realistic
hardening while in vector times shear modulus. assumptions such as
brittle material on size Cross–sectional area of the cut cubical shape of grains
and rate of chip does not change during and participation of all
formation. machining. grains in cutting action
Viscosity effects in water slurry (under the tool tip)
are almost negligible. have been made.
No allowances for grain
size variation. Number
of active grains is
derived assuming slurry
is drawn when tool
recedes.
Rozenberg Brittle fracture Abrasive particles are Involves tedious
et al. incompressible and are of computation and its
(1964) irregular shape but can be solution requires
considered as spheres having numerous integration.
projections whose radii of
curvature are proportional to
the mean dimensions of
particle.
Based on the experimental
evidence, the statistical
distribution of abrasive
particle size d is given by:

(Continued )
Ultrasonic Machining 335

TABLE 2 Continued

Name of Mechanism of material


Investigator removal Assumptions Limitations
  2 3
uðd Þ ¼ 1:095 dNm 1  ddm  1
where N is number of active
abrasive grains, dm is the mean
diameter of grains.
N.H. Cook Hemispherical Abrasive grains are spheres of Model predicts linear
(1966) indentation fracture uniform radius. relationship between
Tool and abrasives are rigid. static stress and MRR,
Viscosity effects are negligible. while MRR drops after a
A linear relationship between certain value of feed
fraction of active grits and force.
ratio of indentation depth to It predicts that MRR is
grit-radius has been assumed. proportional to square
root of the grain radius,
while practically an
optimum value exists.
G.S. Kainth Indentation fracture due Abrasive grains are spherical in Computationally
(1979) to direct hammering shape and follow Rozenberg’s intensive.
action size distribution functions to Predicts linear
take into account particle size relationship between
inhomogeneity. MRR and static force F
Motion of tool remains that is practically not
sinusoidal under loaded true.
conditions. Predicts linear increase in
MRR with grain size,
while and optimum
value exists.
Theoretical machining
rate is higher than
practical values.
Nair and Brittle fracture Abrasive particles are rigid Derivation of the model is
Ghosh spheres computationally
(1985) No consideration for MRR due intensive.
to particle impacting, The volume fractured by a
cavitation or chemical action single abrasive grain is
of slurry to be calculated using
Tool tip motion is SHM and fracture profile.
abrasive particle rests on a
brittle half-space and receives
only a single impact in this
position.
Wang and Combined effect of impact Work-piece is assumed to be Can be used for perfectly
Rajurkar indentation and fracture semi-infinite solid. brittle materials only
(1995) phenomenon Axis of moving grit is like amorphous glass.
perpendicular to the free Results are not true for
surface during machining. materials which exhibit
Speed of abrasive is same as that some plastic behavior
of vibrating tool. like carbides.

(Continued )
336 J. Kumar

TABLE 2 Continued

Name of Mechanism of material


Investigator removal Assumptions Limitations

Lee and Brittle fracture Pre-existing flaws are assumed in Applicable to brittle
Chang the material for the initiation materials only.
(1997) of median or lateral cracks.
Size of median or lateral crack is
related to pseudo pressure
between tool and work-piece.
Cutting tool is assumed to be a
slender column
Wiercigroch Micro-cracking due to MRR is a function of the Applicable to hard and
et al. impacts of grains magnitude of impact force brittle materials only.
(1999) and its frequency. Diamond is Tool geometry changes
uniformly distributed over the with progress in
working part of tool with a machining as the wear
uniform grit size. Ultrasonic on the surface of tool is
vibration amplitude, not uniform.
frequency and tool geometry
remain unchanged.

From all these models, four different mechanisms can be identified,


which are responsible for removal of material from the work surface
(Figure 7). These are described as:

. Material abrasion by direct hammering of the abrasive particles against


the workpiece surface (Cook, 1966; Kainth, 1979; Lee and Chan, 1997;
Miller, 1957; Nair and Ghosh, 1985; Rozenberg et al., 1964; Shaw, 1956).
. Micro-chipping by impact of free moving particles (Miller, 1957;
Rozenberg et al., 1964; Sreejith and Ngoi, 2001).
. Cavitation effect from the abrasive slurry (Shaw, 1956; Soundrajan and
Radhakrishnan, 1986).

FIGURE 7 USM material removal mechanisms (Thoe et al., 1995).


Ultrasonic Machining 337

. Chemical action associated with the fluid employed (Benedict, 1987;


Weller, 1984).

The individual or combined effect of the above listed mechanisms


results in material removal by shear (Shaw, 1956) or by fracture (Cook,
1966; Kainth, 1979; Lee and Chan, 1997; Nair and Ghosh, 1985; Rozenberg
et al., 1964) and by displacement of the material at surface, without
removal, by plastic deformation (Miller, 1957), which occurred simul-
taneously at the transient surface. With porous materials such as graphite,
cavitation erosion has been found to be a significant contributor of
material removal as opposed to hardened steels and ceramics (Riddie,
1973).
Riddie (1973), Bulat (1974) and Willard (1953) reported that the cavi-
tational bubbles formed during the ultrasonic oscillations produce a press-
ure more than 1000 kgf=cm2 on the work surface when they collapse. This
pressure rise also causes material removal. It has been reported that as long
as the working gap between the tool and workpiece surface is more than
the mean size of the abrasive grains used, no significant machining can take
place (Soundrajan and Radhakrishnan, 1986; Wang and Rajurkar, 1995).
The free impact mechanism has been found to be more effective and sig-
nificant with larger grain sizes of abrasive used. With fine grit sizes of abras-
ive grains, the direct hammering mode has been found to be predominant
for material removal.
The initiation and propagation of median as well as lateral cracks has
been considered to contribute greatly to the material removal process in
ultrasonic machining (Ghahramani and Wang, 2001; Deng and Lee,
2002; Halm and Schulz, 1993;) of ceramic composites. The sharp point
of indenter (abrasive grain) produces an inelastic deformation zone and
at some threshold, a deformation-induced flow suddenly develops into a
small crack, termed as a median crack.
An increase in the load causes further growth of the median crack.
Figure 8 shows the mechanism of generation and propagation of cracks
while machining ceramics in USM. During unloading, the median crack
begins to close; inducing the formation of lateral vents (Lee and Chan,
1997; Pei et al., 1995a; Pei et al., 1995c). Upon complete unloading, the lat-
eral vents continue their extension towards the work surface and lead to
chipping.
Zarepour and Yeo (2012) developed a model to predict ductile and
brittle material removal modes while a brittle material is impacted by a
single sharp abrasive particle in micro-ultrasonic machining process. The
model was developed on the basis of indentation fracture theory. The quan-
titative criterion for brittle-ductile transition in material removal were
evolved using threshold kinetic energy consideration; in promoting radial
338 J. Kumar

FIGURE 8 Crack propagation and work surface indentation in USM (Lawn et al., 1980; Marshall,
Lawn, and Evans, 1982).

and lateral cracks. The outcome of this research could be used as a


platform to build reliable models for prediction of MRR based on the
material removal mode.
Sahay et al. (2011) applied Monte Carlo simulation based crystal ball
analysis tool to two popular mathematical models of USM namely Miller’s
model and M.C. Shaw’s model. Effects of input variables such as abrasive
particle size, particle concentration, amplitude of vibration, tool radius
and depth of cut on MRR were analyzed for both models. Simulation of
Shaw’s model indicated that MRR decreases with increase in grain size.
However, literature available on experimental data contradicts this and sug-
gests just the reverse. Miller’s model predicts a steeper relationship between
grain size and MRR. It was concluded that none of the two models
absolutely explains the process; therefore, a realistic modeling possibly is
a combination of the two models or some more factors needed to be taken
into account.

PROCESS CAPABILITIES
USM is valuable process for precision machining of hard, brittle materi-
als. Although USM is not limited by high-hardness materials, best machining
rates are obtained for materials having hardness more than HRC 60
Ultrasonic Machining 339

(Rozenberg, 1973; Saha et al., 1988; Snoyes, 1986; Tsutsumi, 1993; Weller,
1984). Materials such as carbides, ferrites, germanium, ceramics, glass and
tungsten are representative of those that are difficult to process convention-
ally and can benefit the most from the USM process (Balamuth, 1964;
Graff, 1975; Halm and Schulz, 1993; Haun and Schulz, 1994; Hocheng
et al., 1999; Kennedy and Sakaar, 1989; Kremer and Mackie, 1988; Markov,
1977; Spur et al., 1997). When performing drilling operation, USM can pro-
duce holes as small as 76 mm in diameter (Thoe et al., 1998). The best tol-
erance that can be obtained practically in ultrasonic drilling is of the order
of 25 mm; however with special considerations given to slurry circulation
and abrasive selection, tolerances of the order of  10 mm can be achieved
(Kennedy and Grieve, 1975). Holes of up to 64 mm thickness can be drilled
successfully without applying special efforts. Holes as small as 76 mm in
diameter can be drilled, however the depth to diameter ratio is limited to
3:1 (Kennedy and Grieve, 1975; Thoe et al., 1998).
When optimum flushing techniques are used, hole-depth capabilities
can be extended to 150 mm with aspect ratio up to 40:1 (Ghabrial, 1986;
Ghabrial et al., 1982; Koops, 1964; Koval Chenko et al., 1986; Kremer,
1991). However, effective machining rate is reduced for machining of
workpiece thickness more than 12.7 mm, due to inefficient slurry flow
through the cutting gap (Barash and Watanapongse, 1970; Gilmore,
1989; Goetze, 1956; Kaczmarek, 1976). Penetration rates, ranging from
0.025–25 mm=min can be obtained depending upon the shape being
machined, input parameter settings and work material properties
(Markov, 1959; Moreland, 1988a, 1988b; Pentland and Ektermanis,
1965). Surface finish is generally governed by abrasive particle size
(Adithan and Vankatesh, 1976; Dam and Schreiber, 1995; Dvivedi and
Kumar, 2007; Farago, 1980; Guzzo et al., 2003; Hocheng et al., 1999;
Kazantsev, 1963; Komaraiah et al., 1988). From the literature available
on USM, best surface finish has been reported while using 800 grit abra-
sives and is of the order of 0.25 mm (Bhattacharya, 1973).
Because USM is a non-thermal and non-electrical process, the work
material properties remain unaltered (Moreland, 1988a, 1988b; Neppiras,
1972; Nishimura, 1954; Neppiras, 1964; Pandey and Shan, 1980). Table 3
shows the machining performance of USM for different work materials
while using a cold rolled steel tool and boron carbide abrasive (particle size
32 mm) (Mishra, 2005). Table 4 shows a comparison among various
non-traditional machining processes for their shape applications. Table 5
shows the typical values of the machining characteristics of various
processes. A comprehensive review has been carried out on the machining
characteristics of USM, by comparing and analyzing the investigations of
many researchers. Table 6 presents detailed review on machining character-
istics and process capabilities of USM.
340 J. Kumar

TABLE 3 Machining Performance of USM (Benedict, 1987; Mishra, 2005)


Max. machining Avg. machining rate
Material Ratio of MRR to TWR area (sq. mm) (mm=min)

Glass 100:1 25.8 3.81


Ceramics 75:1 19.4 1.51
Germanium 100:1 22.6 2.16
Tungsten carbide 15:1 7.7 0.25
Tool steel 1:1 5.6 0.13
Mother of pearl 100:1 25.8 3.81
Synthetic Ruby 2:1 5.6 0.51
Carbon-graphite 100:1 19.4 2.00
Ferrite 100:1 22.6 3.18
Quartz 50:1 19.4 1.65
Boron Carbide 2:1 5.6 0.20
Glass-bonded mica 100:1 22.6 3.18
Materials successfully
machined using USM
Aluminium Cold rolled steel Micarata Ti-alloys
Aluminium Oxide Febony Molybdenum Tungsten
Barium Titanate Ferrite Mother of Pearl Tungsten carbide
Beryllium Oxide Garnet Plaster of Paris Thorium Oxide
Boron carbide Germanium Quartz Uranium Carbide
Boron composites Glass Ruby Zirconium Oxide
Brass Glass bonded Sapphire Diamond
Calcium Mica Silicon Carbide Fiber-reinforced
Carbides Graphite Silicon Nitride Plastics
Carbon Hardened Steel Stainless Steel High Pressure
Ceramics Limestone Stellite Laminates
Composites Lithium-fluoride Tool Steel

TABLE 4 A Comparison of Shape Applications of USM to Various Processes (Singh, 2007)


Holes
Through
Precision small holes Standard Through Cavities Surfacing Cutting

Dia Dia Length Length Double Con-


Process <.025 mm >.025 mm <20 mm <20 mm Precision Standard touring Shallow Deep

USM ___ ___ Good Poor Good Good Poor Poor ___
AJM ___ ___ Fair Poor Poor Fair ____ Good ___
ECM ___ ___ Good Good Fair Good Good Good Good
CHM Fair Fair ____ ____ Poor Fair ____ Good ___
EDM ___ ___ Good Fair Good Good Fair Poor ___
LBM Good Good Fair Poor Poor Poor ____ Good Fair
PAM ___ ___ Good ___ Poor Poor ____ Good Good

Abbreviations: USM – ultrasonic machining; AJM – abrasive jet machining; ECM – electro-chemical
machining; CHM – chemical machining; EDM – electric discharge machining; LBM – laser beam
machining; PAM – plasma arc machining.
Ultrasonic Machining 341

TABLE 5 Typical Values of the Machining Characteristics of Various Processes (Singh, 2007)
MRR Depth of Surface
Process (mm3=min) Tolerance (mm) Surface (mm) CLA damage (mm) Power (watts)

USM 300 15 0.2–0.5 25 2400


AJM 0.8 50 0.5–1.2 2.5 250
ECM 15,000 50 0.1–2.5 5.0 100,000
CHM 15 50 0.5–2.5 50 –
EDM 800 15 0.2–1.2 125 2700
EBM 1.6 25 0.5–2.5 250 150 (avg.)
LBM 0.1 25 0.5–1.2 125 2000 (peak)
PAM 75,000 125 Rough 500 50,000
CM 50,000 50 0.5–5.0 25 3000

Abbreviations: USM – ultrasonic machining; AJM – abrasive jet machining; ECM – electro-chemical
machining; EDM – electric discharge machining; EBM – electron beam machining; LBM – laser beam
machining; PAM – plasma arc machining; CHM – chemical machining; CM – conventional machining.

OPERATING CHARACTERSITICS
Material Removal Rate
To identify the potential factors affecting material removal rate in USM,
a cause and effect diagram was constructed (Figure 9). As the diagram
indicates, the material removal rate in USM is dependant on four primary
factors workpiece; tool; slurry and machine related factors. The literature
corresponding to these factors has been extensively reviewed and presented.

Workpiece Properties
Komaraiah and Reddy (1993a) investigated the influence of work material
properties such as fracture toughness and hardness on material removal rate in
ultrasonic machining of hard and brittle materials. MRR was reported to
decrease with an increase in work material hardness and fracture toughness
in almost linear fashion under controlled experimental conditions. In another
investigation by Komaraiah et al. (1988), the MRR was reported to depend upon
the brittleness ratio (ratio of work hardness to elastic modulus) of the work
material. Deng and Lee (2002) reported the MRR to be low while machining
composites of higher fracture toughness such as whisker-reinforced compo-
sites. The particle reinforced composites yielded higher values of MRR on
account of their low fracture toughness. The composites of higher flexural
strength demonstrated better surface integrity while machining with USM.
Guzzo et al. (2003) outlined the ultrasonic abrasion of different hard
and brittle materials using stationary USM. Results show that machining
rate decreased with increase in hardness of the work material. Majeed
et al. (2008) outlined the machining of Al2O3=LaPO4 composites using sta-
tic USM. Results show that an increase in the hardness of the composite
TABLE 6 Review on Machining Characteristics of USM

342
Characteristics
Investigator Work materials Process Conditions investigated Results

Dam et al. (1995) Glass, SiC Fixed: MRR MRR=TWR Dev SR (mm)
Al2O3, TiB2 Tool: Steel TWR (mm).
Hot pressed Slurry: B4C Deviation in Glass: 108 10 13.5
silicon nitride Grit: 280 hole size Al2O3: 30.8 25 9.0
(HPSN) Variables: (Dev.) TiB2 : 23.5 100 3.2
Zirconium Oxide work material Surface SiC: 4.5 75 6.0
(TZ3YB) roughness HPSN: 1.2 75 4.0
(SR) TZ3YB 0.75 150 2.0
Adithan and Glass Fixed: Surface Load (kg) SR (mm) Conicity Roundness
Krishnamurthy Tool: MS roughness (mm) (mm)
(1978) Slurry: B4C (SR) 0.15 10.5 75 55,55
Grit: 280 Conicity, 0.35 12.5 65 53, 68
Variable: Out-of- 0.50 17.0 85 40,43
Static load Roundness 0.85 18.2 70 45,75
(at entry, exit) 1.00 18.6 40 43,117
Hocheng and Hsu Fiber reinforced Fixed: Surface Surface roughness: 1.2–2.0 mm
(1995) Plastics Tool: MS Roughness Hole clearance: 0.12–0.27 mm
Slurry: SiC Deviation in
Variable: hole size
Grit: 150–600 (Dev.)
Conc: 13–26%
Majeed et al. Al2O3=LaP04 Fixed: MRR Hardness MRR (mm3=s)
(2008) Composites Slurry: B4C (kg=mm2)
Grit: 280 Tool 1 Tool 2
Tool: L.C.S. 400 0.65 0.25
Variable: 800 0.60 0.20
Work Hardness 1200 0.45 0.12
Tool geometry 1600 0.10 0.05
(tool 1: hollow, 2: solid)
Guzzo et al. (2003) Quartz crystal Fixed: TWR TWR: 3.4–16.5 mm=s
Slurry: SiC Surface SR: 1.0–1.9 mm
Tool: S.S. roughness
Variable: (SR)
Grain size
6–50 mm
Dvivedi and Titanium alloy Fixed: Slurry Surface SR: 0.96–3.40 mm
Kumar (2007) (6Al-4 V) (B4C) Roughness Optimized value of SR: 0.97 mm
Variable: (SR)
Tool material
Grain size:
18–64 mm
Slurry Conc.
Komaraiah et al. Glass (G) Fixed: Surface Grit Size A P F G
(1988) Ferrite (F) Static load Roughness SR (mm)
Porcelain (P) Amplitude (SR) 220 1.2 3.9 1.8 2.6
Alumina (A) Slurry: SiC MRR 280 0.7 2.1 0.9 1.6
Variable: Out-of- 320 0.4 1.4 0.4 1.3
Grit size roundness MRR (mm3=min)
220, 280, 320 220 1.8 20.5 2.2 4.5
Out-of-roundness
(mm)
180 45 120 56 82
Jadoun et al. Ceramic Fixed: Tool wear rate TWR : 0.072–0.34 mm3=min
(2006) Composites Slurry: SiC (TWR) Optimized value: 0.075 mm3=min
Static load
Variable:
Tool materials:
HCS, WC, HSS
Grit: 18–64 mm
Power: 40–60%
Komaraiah and Glass Fixed: MRR (mm3= Tool MRR Tool H
Reddy (1993b) Slurry and grit size (SiC= min) wear
220) Tool wear MS 3.0 0.77 250
Variable: Tool (mm) Ti 3.4 0.56 255
MS, Ti, S.S., Tool hardness S.S. 4.2 0.22 375
Niamonic-80, (H) after Ni-80 5.8 0.11 590
Silver steel machining Si Steel 5.5 0.20 535
(VPN) (Depth of
machining:
24 mm)

(Continued )

343
344
TABLE 6 Continued

Characteristics
Investigator Work materials Process Conditions investigated Results

Deng and Lee Ceramic Fixed: MRR Fracture MRR SR


(2002) Composites Slurry: B4C=80 Surface Toughness(mm3= (mm)
(alumina based) Static load: 10 N Roughness min)
Variable: (SR) 5.0 0.95 1.75
Fracture toughness 6.0 0.75 1.20
(work) 7.8 0.44 0.82
8.6 0.38 0.60
2
Komaraiah and Glass, ferrite Variable: MRR MRR / K 1:5PH0:5 : Nf
Reddy (1993a) Alumina, WC Work hardness P is the static load,
Porcelain Fracture toughness f is frequency
K is fracture toughness,
H is hardness
N is effective no.
of grains in the gap
Zhang et al. Alumina ceramics Tool: L.C.S. MRR (mm3= Load (N): 10 15 20 25 30
(1999) B4C=120 min) MRR: 3.5 4.5 5.5 5.0 3.0
Variable: Grit (mm) 50 100 150 200
Static load MRR: 4.0 5.3 5.5 5.3
Amplitude Amp (mm) 03 06 09 12
Grain size MRR: 2.8 4.5 5.6 5.8
Ultrasonic Machining 345

FIGURE 9 Cause-and-effect diagram for MRR, TWR in USM.

(due to addition of LaPO4) improves the machining rate up to a critical


limit after which it tends to stabilize. The use of hollow tools was also
reported to improve the MRR obtained.
Kumar et al. (2009) investigated the influence of work material
properties on the performance indices of ultrasonic machining. Effects
of input parameters namely abrasive material type, grain size, tool geometry
were observed on material removal rate, penetration rate and tool wear rate
for six different work materials and using three different tool materials.
The materials with higher fracture toughness were found to take longer
times for machining with USM while the hard and brittle materials gave
better results in terms of the machining efficiency. Work materials with
higher hardness and toughness caused more tool wear whereas the softer
and brittle work materials resulted in lesser tool wear.
Khairy (1990) presented an assessment of the influence of work material
properties on mechanism of material removal in USM. Results showed that
hard and brittle materials such as glass were machined by brittle fracture at
selective cleavage planes, whereas the tough materials face considerable
plastic deformation before failure. Dam, Quist and Schreiber (1995) inves-
tigated the material removal rate in ultrasonic drilling of several different
ceramic materials. Results showed that for tougher work materials, the
MRR observed was quite low as compared to the hard and brittle materials.
It has been concluded that productivity by USM (in terms of machining
rate) is primary determined by the brittleness of the work material
(Kazantsev, 1966; Markov, 1959; Neppiras, 1964). The plasticity of work
material is associated with low productivity. The impact hardness has been
found to have an adverse effect on machining rate (Komaraiah et al.,
346 J. Kumar

1988). However, while machining annealed steel, the machining rates


observed have been found to be significantly better than normalized or
quenched ones (Kubota et al., 1977). Kennedy and Grieve (1975)
concluded that such a behavior contradicted the generally accepted
criterion. However, this unusual behavior was explained better by
Kremer (1981), considering the increase in tensile strength of steel upon nor-
malizing or quenching, which is unfavorable from the point of view of USM.

Tool Characteristics
Komaraiah and Reddy (1993b) investigated the influence of tool
material properties i.e., hardness on the material removal rate in USM of
glass. Results showed that the MRR increased with an increase in the hard-
ness of the tool material. The different tool materials were arranged in the
increasing order of superiority as mild steel < titanium < stainless steel <
silver steel < niamonic-80 A < thoriated tungsten. The tool materials used
were found to undergo a significantly different amount of work-hardening,
which contributed to the variation in their machining performance.
Neppiras (1957) using other tool materials gave the following ranking:
copper < stainless steel < silver steel < mild steel < brass < tungsten carbide.
Kumar et al. (2008) reported achievement of higher material removal rates
using a high carbon steel tool, which had higher hardness in comparison to
other tools used for the experimentation. Tools with diamond tips have
been shown to have good material removal characteristics (Kazantsev, 1966).
It has been reported that the machining rate is directly proportional to
the tool form (Kennedy and Grieve, 1975; Farago, 1980; Neppiras, 1972)
and shape factor (ratio of tool perimeter to tool area). The tool form defines
the resistance to slurry circulation: a tool of narrow rectangular cross-section
yielding a better machining rate than one with a square cross- section of the
same area (Kennedy and Grieve, 1975; Mcgeough, 1988; Snoyes, 1986). Use
of hollow tools has been reported to result in higher rates of material
removal than ones with solid geometry for same area of cross-section (Pan-
dey and Shan, 1980). Goetze (1956) presented a study on effect of tool
geometry on the penetration rate obtained in the USM of ketos tool steel.
For tools with equal contact areas, an increase in the penetration rate was
reported for tools with larger perimeters. This was explained on the basis
of difficulty of adequately distributing the abrasive slurry over the machin-
ing zone (Kremer, 1981; Neppiras, 1957; Sreejith and Ngoi, 2001).
Several authors have reviewed the theory and art of designing the tool=
horns, but it is not yet fully understood (Amin et al., 1993; Balamuth, 1964;
Dam and Jensen, 1993; Frederick, 1965; Kaczmarek, 1976; Kumehara,
1984; Merkulov, 1957; Satyanaryana and Krishnan Reddy, 1984). Detailed
Ultrasonic Machining 347

guidelines for tool design for optimum MRR have been described by
Rozenberg et al. (1964). Traditional methods of acoustic horn design are
based upon differential equation which considers the equilibrium of an
infinitesimal element under the action of elastic and inertia factors, which
is then integrated over the horn length to achieve resonance (Pandey and
Shan, 1980). Typical design includes cylindrical, stepped, conical and expo-
nential types. Recently, finite element modeling (FEM) has been used
(Benedict, 1987; Seah et al., 1993) to design symmetric horn shapes. The
analysis can take into the consideration the weight of the tool (Benedict,
1987). Dam and Jensen (1993) has claimed that a horn can be designed
that converts the longitudinal ultrasonic action into a mixed lateral and
longitudinal vibration mode. This lateral motion obviously aids contouring
work (Kremer, 1991).

Slurry Properties
Various investigators (Anantha Ramu and Krishnamurthy, 1989; Barash
and Watanapongse, 1970; Khairy, 1990; Koops, 1964; Koval Chenko, 1986;
Pentland and Ektermanis, 1965; Scab, 1990; Willard, 1953) have reported
results indicating that the rate of material removal for a certain abrasive
is a function of its concentration, grain size and hardness besides the feed
system.
On increasing the abrasive grit size or slurry concentration, an opti-
mum value of MRR is reached. Any further increase in either aspect results
in difficulty in the larger grains reaching the cutting zone (Goetze, 1956;
Koops, 1964) or a subsequent fall in MRR. Guzzo et al. (2003) reported
a substantial increase in MRR obtained while using abrasive of larger grain
size on account of the increase in the stress caused by the impact of abrasive
particle over the workpiece surface. Neppiras (1964) and Markov (1977)
reported that when grain size is comparable to the amplitude of vibration,
the optimum level of MRR can be reached. Experimentally the ratio of the
double amplitude to the mean size of the principal fraction of abrasive is
0.6 to 0.8. Goetze (1956) has reported the optimum value of slurry concen-
tration to be close to 12% for all the abrasive grit sizes used in the investi-
gation. The optimum concentration is thought to be one providing a single
layer of abrasive over the entire work surface (Kremer, 1981). The values
given for the optimum concentration are inconsistent, with a range of
30% to 60% (Markov, 1959), 25% to 40% (Neppiras, 1972) and 15% to
40% (Nishimura, 1954).
The disagreement in the quoted values for the optimum range of con-
centration can be attributed to the variation of concentration within the
working zone under the tool. Obviously such effective local concentration
348 J. Kumar

under the tool can not be the same as that of the feed suspension especially
when the static load is too high for the particular acoustic setting (Kremer,
1991). Kennedy and Sakaar (1989) pointed out the difficulty of machining
a flat at the bottom of a hole because of uneven slurry distribution across
the machining face, resulting in fewer active grits at the tool centre. Kazant-
sev (1963, 1966) claimed that the forced delivery of the slurry increased the
output of USM five fold without the need to increase grit size or machine
power. When compared with suction pumping system, it yielded a 2–3 times
higher MRR. Pentland and Ektermanis (1965) and others (Koops, 1964,
Koval Chenko et al., 1986; Scab, 1990) found that by improving slurry cir-
culation, the adverse effects such as contamination and blockage can be
reduced or overcome. Barash and Watanapongse (1970) reported a four-
to five-fold increase in MRR while increasing the fluid pressure from 10
to 90 psi, attributed to suppression of cavitation at higher pressure of the
slurry fluid.
The hardness of the abrasive has been found to affect the machining
rate. For machining of soda glass, the removal rate with boron carbide
has been reported to be 15–20% higher than that with silicon carbide
(Kennedy and Grieve, 1975). Ramulu (2005) reported that use of boron
carbide abrasive resulted in material removal rates which were approxi-
mately 75% higher than the silicon carbide abrasive for the 400 grit size
and 320% higher for 220 grit size, while machining silicon carbide cer-
amics. In general, use of a harder abrasive (such as boron carbide) yields
best values of machining rate, provided that other factors remain
unchanged. The effect of slurry hardness on MRR has been found to be
dependant on the other experimental conditions such as work material
properties, tool properties, amplitude of vibration and static load; which
could be regarded as the reason for a wide inconsistency of the results
reported in the literature available on USM, e.g., for a particular material,
increasing the slurry hardness by 50% may yield an improvement of the
order of 25% in the MRR, but for some other material, the increment in
MRR might be 10%.

Operating Parameters
Power primarily determines the mass of the tool-horn assembly that can
be utilized for an application and the frontal cutting area of the tool. The
more is power available in an ultrasonic machine, the larger the frontal cut-
ting area of the tool can be supported (Benedict, 1987). The amplitude of
vibration (n) has been found to affect the machining performance of USM
by a number of investigators (Anantha Ramu and Krishnamurthy, 1989;
Goetze, 1956; Kainth, 1979; Kazantsev, 1963; Miller, 1957; Neppiras, 1957;
Ultrasonic Machining 349

Rozenberg, 1973; Weller, 1984; Zhang et al., 1999). Higher amplitude is


obtained by using a tool with a larger transformation ratio i.e., the ratio
of transducer=tool diameter (Kennedy and Grieve, 1975; Kremer, 1991;
Kremer et al., 1983, Thoe et al., 1998). Shaw (1956) showed that MRR is
proportional to n3=4 while other researchers (Goetze, 1956; Kainth, 1979;
Kazantsev, 1966; Miller, 1957; Thoe et al., 1995) have advocated that
MRR is linearly proportional to n and yet others (Neppiras, 1957; Neppiras,
1964; Pentland, 1965; Rozenberg et al., 1964; Wang and Rajurkar, 1995)
have suggested that MRR depends upon n2 for constant frequency and
static load conditions. Goetze (1956) reported a linear increase in MRR
while increasing the amplitude of vibration, provided all other factors such
as frequency of vibration and abrasive grit size are kept constant. The linear
trend of MRR while increasing the amplitude has been found to be more
prevalent when high impact strength materials are machined or fine abras-
ive powders are used for machining. Lee and Chan (1997) reported an
optimum value of amplitude beyond which the MRR obtained tends to
stabilize. When a larger grit size of the abrasive is coupled with a low value
of amplitude, the MRR obtained is reported to be substantially poor due to
ineffective circulation of slurry under the tool.
Experiments conducted by Neppiras (1972) have shown that in the
range of 20 to 50 kHz, the removal rate is proportional to square root of
the vibration frequency. However, Kazantsev (1966) stated that the abrasion
rate is proportional to the frequency while the non linear frequency depen-
dence of machining rate is due to the variation in abrasive concentration in
the working zone. Some other researchers have also reported the linear
dependence of machining rate on frequency of vibration (Gilmore, 1989;
Goetze, 1956; Zhang et al., 1999). Above an upper threshold value, the
MRR falls off rapidly where MRR is proportional to square root of fre-
quency (Kazantsev, 1963, Nishimura, 1954).
Kainth (1979) carried out an analysis considering the non-uniformity
of abrasive grains to assess the relation between the removal rate and sta-
tic load=amplitude. Their calculations yielded approximately a linear
relation between material removal rate and static load. Rozenberg
(1964) has also reported similar results while all other factors are kept
unchanged. Koops (1964) indicated that use of a smaller than optimum
value for static load is better for reducing the abrasive wear and increas-
ing the tool life. Above the optimum value, the MRR decreases owing to a
reduction in the size of abrasive grains reaching the cutting interface
(due to crushing effect) and inefficient slurry circulation (Graff, 1975;
Kennedy and Grieve, 1975; Neppiras, 1972; Rozenberg, 1973; Snoyes,
1986; Zhang et al., 1999).
The optimum static load for the maximum machining rate has been
found to be dependant on the tool configuration, amplitude of vibration
350 J. Kumar

and mean grit size (Gilmore, 1991; Kremer, 1991; Neppiras, 1964). A
previous analysis carried out by Markov (1959) on the experimental results
of ENIMS yielded a wide range for the power exponent relating MRR with
amplitude for different static loads. The values ranged from 0.5 to 1.7
depending upon the value of static load used.
Lalchnuanvela et al. (2012) reported an experimental investigation on
high alumina ceramics based on central composite second-order rotatable
design. In this study, abrasive grit size, slurry concentration, power rating,
tool feed rate, slurry flow rate are considered as input variables. The opti-
mal parametric combinations of input variables for maximum MRR and
minimum value of roughness have been obtained through multi-objective
optimization.
Kumar and Khamba (2010a) investigated the material removal rate in
ultrasonic machining of pure titanium using Taguchi’s method (L-18
OA) for experimental design. Effects of input parameters such as tool
material, abrasive, grain size and power rating of USM were assessed on
MRR using ANOVA test. All the factors considered were found to be signifi-
cant at 95% level. The modeling of MRR using dimensional analysis
method was also performed.
Gauri et al. (2011) analyzed two sets of past experimental data on
USM process using three different methods dealing with multiple corre-
lated responses (MRR, tool wear and Surface roughness) and the optimi-
zation performances of the three methods were subsequently compared.
It was reported that both the weighted principle component (WPC)
and principal component analysis (PCA)-based TOPSIS methods result
in a better performance than PCA-based grey relational analysis.
The WPC method was found to be best because of simpler computational
procedure.
Rao et al. (2010) performed optimization of MRR as response subjected
to constraint of surface roughness. The process variables considered for
optimization were—amplitude of vibration, frequency, and mean diameter
of particles, static feed force and slurry concentration. The optimization
was performed using three optimization algorithms namely artificial bee
colony, harmony search and particle swam optimization. The results
showed that three methods outperformed the genetic algorithm with a
considerable improvement.
Singh and Gill (2009) presented the design of fuzzy logic based model
to simulate MRR in ultrasonic drilling of porcelain ceramic. The model can
be referred by machine operator from time to time and can also be used in
process planning by practicing engineers. The model was based on two
input signals—namely the depth of cut and time of cut. The application
of Chi-square test indicated that the values of MRR predicted by model
Ultrasonic Machining 351

were well in agreement with experimental values at 0.1% level of


significance.

Tool Wear in USM


Tool wear is an important variable influencing both MRR and accu-
racy in USM (Adithan, 1981; Goetze, 1956; Laroiya and Adithan, 1993;
Venkatesh, 1983). Tool wear in USM has been classified into two types:
longitudinal wear, WL (Markov, 1977; Seah et al., 1993) and lateral=
diametral wear, WD (Adithan and Venkatesh, 1978) as depicted in
Figure 10. Besides these, some of the tool wears occur as a result of cavi-
tation or suction wear (Riddie, 1973; Seah et al., 1993; Willard, 1953).
Tool wear in USM is a complex phenomenon and is affected by a number
of factors such as static load; work material; tool material; tool size; type
of abrasive and its grain size; machining time; depth of hole drilled, etc.
As a result of the tool wear, both tool length and weight decrease, which
affects the resonance frequency of the machine and reduces the ampli-
tude of vibration (Adithan, 1981; Adithan and Venkatesh, 1978; Adithan
and Krishnamurthy, 1978; Adithan and Venkatesh, 1976), thus, lowering
the MRR.
Adithan and Venkatesh (1974) reported that the tool wear is
maximum at a particular static load, which may be considered optimum
for the point of view of MRR. The tool wear increases linearly with the
total depth of holes drilled. As the depth of hole drilled increases, there
is a reduction in the MRR and an increase in the associated tool wear
(Adithan and Venkatesh, 1976). The tool wear is proportional to the cut-
ting time and the rate of tool wear has been found to increase with time
(Komaraiah and Reddy, 1993b; Soundrajan and Radhakrishnan, 1986).
Adithan (1981) found that for all tool and work combinations, the tool
wear rate is proportional to the wear that has already taken place. The tool
wear tends to increase when harder and coarser abrasives are used. As a
consequence, harder abrasives like boron carbide cause higher tool wear

FIGURE 10 Tool wear in USM (Adithan, 1981).


352 J. Kumar

compared to softer abrasives like silicon carbide for a tool of the same
cross-sectional area (Adithan and Venkatesh, 1974; Venkatesh, 1983).
Anantha Ramu and Krishnamurthy (1989) reported an optimum slurry
concentration range (1: 5.7) from tool wear aspect in USM of transform-
ation toughened ceramics.
Dam et al. (1995) has reported that work material properties such as
hardness and toughness affect the tool wear in ultrasonic machining. Their
results showed that work materials with higher fracture toughness and
hardness cause more tool wear. Anantha Ramu and Krishnamurthy
(1989) have reported that transformation toughened ceramics show poorer
USM behavior giving relatively high tool wear. This has been attributed to
their higher fracture toughness which is unfavorable from the point of view
of tool wear in USM. Adithan (1981) reported absence of lateral wear in
machining of softer materials such as porcelain, while it was found to be
more prevalent in case of hard work materials such as tungsten carbide.
Adithan (1981) has reported that stainless steel tools exhibit low tool
wear as compared to tungsten carbide or mild steel tools. This is due to
high resistance to cavitation erosion of stainless steel. In USM, hardness
of the tool increases by work hardening, therefore the penetration of the
abrasive grains into the tool decreases resulting in higher MRR (Komaraiah
and Reddy, 1993b). In addition, material removal from the periphery of the
work zone becomes greater so that a convex surface is formed in the work-
piece. This causes plastic deformation of the centre of the tool face, form-
ing a dish. It has also been found that the degree of hardening is higher at
the periphery and lowest at the centre for all tool materials (Komaraiah
and Reddy, 1993a, 1993b). As a result, soft materials such as copper and
brass are unsuitable as tools as they develop burrs at large oscillatory ampli-
tudes (Neppiras, 1964). They are also acoustically poor and attenuate the
stress wave in large tools. The use of hard metals such as tungsten carbide
reduces plastic deformation of the tool surfaces (Halm and Schulz, 1993).
To decrease longitudinal wear (WL), use of a tool material with a high
value of product of hardness and impact strength (e.g., Niamonic-80 A) has
been recommended by Komaraiah and Reddy (1993b). However, the
lateral wear (WD) has been found to be independent of the impact strength
(Adithan, 1981; Komaraiah and Reddy, 1993b). Komaraiah and Reddy
(1993a) ranked the various tool materials in the order of superiority
(decreasing tool wear) as mild steel > titanium > stainless steel > silver
steel > thoriated tungsten > niamonic-80 A.
Kumar et al. (2008) compared the machining performance of high car-
bon steel and titanium alloy tools in USM of titanium. High carbon steel
tool was found to experience more tool wear as compared to titanium alloy
tool because of its higher hardness and poor toughness and impact
strength as compared to titanium alloy tool. Kumar et al. (2009) reported
Ultrasonic Machining 353

an experimental study on investigation of tool wear rate in ultrasonic


machining of pure titanium (ASTM Grade-I) using Taguchi’s method.
Power rating and tool material were identified as the input parameters
contributing most to the variation in TWR. Optimal results for tool-wear
rate were obtained with the use of a tool material with optimum hardness
and toughness (titanium alloy), a softer abrasive (alumina), fine grit size of
the slurry and low power rating of the machine.
Jadoun et al. (2006) reported that use of a tool with higher abrasion
resistance such as tungsten carbide results is lesser tool wear rate as com-
pared to high carbon steel or high speed steel tools. Similar results have
been reported by Markov (1959, 1977).

Surface Quality and Hole Integrity


USM does not generate significant heating which might otherwise lead
to the development of a thermally damaged layer=zone or residual stress.
Abrasive grain size has been found to be the main factor governing the
workpiece accuracy and surface finish (Benedict, 1987; Dvivedi and Kumar,
2007; Goetze, 1956; Guzzo et al., 2003; Kennedy and Grieve, 1975; Komar-
aiah et al., 1988; Kumar et al. (2008), Thoe et al., 1998). Neppiras (1964)
concluded that surface roughness is closely controlled by abrasive grain size
alone. Kumar et al. (2008) and others (Guzzo et al., 2003) have also
reported similar results.
A decrease in abrasive grain size while machining with USM leads to
lower values of surface roughness; in addition, the accuracy of the hole
drilled is improved (Adithan and Venkatesh, 1976; Benedict, 1987;
Komaraiah et al., 1988). A better surface finish is also obtained on the bot-
tom walls of the cavity (Mishra, 2005; Moore, 1985; Moreland, 1988a; Scab,
1990; Smith, 1973). Dam et al. (1995) reported that better surface finish is
obtained when feed rates and depth of cut are increased. Komaraiah and
Reddy (1993b) compared the performance of stainless steel, titanium
and niamonic-80 tools for surface finish of the glass workpiece while
machining with USM. Results showed that tool materials with higher wear
resistance (nimonic-80) gave better surface finish as they retain their shape
and finish even under the repeated impact of abrasive particles.
The production inaccuracies in USM consist of both dimensional inac-
curacy (oversize) and form inaccuracy (conicity) (Adithan and Venkatesh,
1976; Benedict, 1987; Komaraiah et al., 1988; Markov, 1959; Rozenberg,
1973). Conicity can be reduced by using tungsten carbide and stainless
steel tools (Adithan and Venkatesh, 1976), an internal slurry delivery sys-
tem (Kremer, 1981) and tools with negative tapering walls or fine abrasives
(Kennedy and Grieve, 1975; Kremer, 1981; McGeough, 1988; Neppiras,
354 J. Kumar

1964). Adithan and Venkatesh (1974) reported that oversize is greatest at


entry and increase in diameter-length ratio increases lateral vibrations
which cause greater oversize.
Shaw (1956) and Komaraiah et al. (1988) has shown that surface rough-
ness improves with increase in static load which reduces the abrasive size
and suppresses the lateral vibrations of the tool, so minimizing the pro-
duction inaccuracies in the hole drilled. Komaraiah et al. (1988) reported
that the out-of-roundness of the drilled holes increases with a correspond-
ing increase in the ratio of work material hardness to Young’s modulus of
elasticity. Adithan and Venkatesh (1976) found that the oversize with rec-
tangular tools was greater than that obtained with circular tools.
It was established by Komaraiah et al. (1988) that workpiece materials
with higher ratio of hardness to elastic modulus involve inferior surface
quality. The materials which observed higher MRR were also reported to
have higher surface roughness values. Figure 11 shows a comparison of
the surface quality obtained (under varied conditions of H=E ratio) while
machining different materials with USM and RUM (Komaraiah et al.,
1988). Gilmore (1989) reported that slightly better surface finish is
obtained while machining a hard material such as ceramics as compared
to that obtained in machining of the material with lower hardness. Markov
(1977) suggested that the relative roughness for two different materials can
be related to their respective volumetric machining ratio in a square
relation.

FIGURE 11 Effect of Hardness (H) to Elastic Modulus (E) ratio on surface roughness in USM and
RUM (Komaraiah et al., 1988).
Ultrasonic Machining 355

Dam et al. (1995) concluded that the work materials can be graduated
according to their respective machining rates, so that the most productive
materials give the greatest surface roughness and vice-versa. Therefore,
higher productivity is not obtained without cost as the surface roughness
increases. Also, the deviation of the diameter of drilled holes from the
nominal diameter (10 mm) has been found to be related with the material
properties such as toughness and hardness. The materials with high hard-
ness have exhibited less deviation in the hole size, whereas the materials
with more toughness have exhibited larger deviation (Figure 12).
Kremer and Mackie (1988) reported that USM of graphite resulted in
poor surface finish due to cavitation, contamination and debris blockage.
Adithan and Venkatesh (1976) found that for the same abrasive size and
static load, the surface roughness for glass workpiece was almost double
that for graphite. Anantha Ramu and Krishnamurthy (1989) compared
the surface quality obtained while machining alumina-based and zirconia-
based ceramics with USM. Results show that better surface finish was obtained
for alumina-based ceramics and the oversize for the drilled hole was also lesser.
Singh and Khamba (2003a, 2003b, 2006, 2007a, 2007b, 2008) investi-
gated the machining characteristics of titanium alloy (Ti-6Al-4V) using
stationary USM. Results show that optimum MRR and TWR was achieved
with boron carbide as abrasive material with grit size 220 and stainless steel
as tool material. Optimum surface quality was generated while using a stain-
less steel tool; slurry concentration of 25% and slurry temperature equal to
27 C. The surface finish has been reported to be better (50 microns) than
that obtained while machining brittle materials such as ceramics. Dvivedi
and Kumar (2007) investigated surface quality in USM of titanium alloy

FIGURE 12 Deviation of drilled holes from nominal diameter (10 mm) in USM (Dam et al., 1995).
356 J. Kumar

(Ti-6Al-4 V). Results show that the best results for surface quality were
obtained with H.C.S. tool; medium grit size (320) and low power rating
of USM machine (40%). The other factors such as slurry concentration
were found to be relatively insignificant from the point of view of surface
quality. Figure 13 shows the modes of material removal while a ductile
and tough material such as titanium is machined using USM. It could be
well observed that the ductile failure of machined surface has taken place
(Figure 13a) under low energy input condition. Whereas under high
energy input conditions, a mixed mode of material removal (dominated
by brittle fracture) could be seen (Figure 13b).
Nath et al. (2012) investigated the effect of micro-chipping on the hole
integrity (such as entrance chipping, wall roughness and sub-surface dam-
age). The material removal mechanism in the lateral gap was also studied.
This was achieved by conducting experiments on three advanced structural
ceramics, namely silicon carbide, zirconia and alumina. It was established
that the sliding and rolling (abrasion) mechanisms by larger abrasives take
part in material removal, however they also produced micro-cracks in radial
direction resulting in sub-surface damage and higher wall roughness.
Figure 14 shows the schematic of the edge chipping of a resultant hole,
as proposed by Nath et al. (2012).
Baek et al. (2013) focused their research on enhancement of the sur-
face quality in USM of glass using a sacrificing coating (of hard wax) on
the substrate (glass). After drilling the holes with USM, the coating was
removed by cleaning. The wax coating protected the substrate from devel-
oping the cracks. It was also reported that the coating could eliminate the
out-of-roundness of the drilled hole, particularly at the exit section, which
otherwise is impossible to contain, even by using worn-out tools.
Majeed et al. (2008) performed research on A.E. monitoring of ultra-
sonic machining of Al2O3=LaPO4 composites. The effect of LaPO4 content
on machining was studied by analyzing the acoustic emission signals

FIGURE 13 Material removal modes while machining pure titanium with USM (Kumar and Khamba,
2010a).
Ultrasonic Machining 357

FIGURE 14 Schematic diagram of edge chipping of a resultant hole in USM (Nath et al., 2012). (Figure
available in color online.)

emitted by the workpiece during machining. It was concluded that the


machinability of alumina can be enhanced by dispersion strengthening
through the addition of LaPO4. There was found to be a critical content
(30% by weight) of LaPO4 for achieving best results. The observed
reduction in MRR with higher-order LaPO4 was attributed to the lack of
sinterability with high LaPO4 content.

ROTARY ULTRASONIC MACHINING


Churi et al. (2009) investigated the machining characteristics of
titanium alloy using rotary ultrasonic machining. Cutting force, material
removal rate and surface roughness were investigated using different
machining variables (ultrasonic power, spindle speed and feed rate). All
the three parameters investigated were found to be significant for their
effect on the machining characteristics. Cong et al. (2012a) reported a
study on power consumption in rotary USM of carbon fiber-reinforced
plastic (CFRP) composites.
They reported an experimental investigation on the effects of input
variables such as ultrasonic power, tool rotational speed, feed rate and type
of CFRP material on power consumption of elements of RUM setup such as
ultrasonic power supply, spindle motor, coolant pump and entire RUM sys-
tem. Results indicated that the type of CFRP material significantly affected
power consumption of ultrasonic power supply and spindle motor, whereas
it did not have much effect on entire RUM system. The increase in
358 J. Kumar

ultrasonic power or tool rotation speed resulted in increase of power


consumption percentage of spindle motor and coolant pump.
Liu et al. (2012b) proposed a mechanistic model for cutting force in
RUM of brittle materials. The relationship between cutting force and input
variables (such as spindle speed, feed rate, ultrasonic vibration amplitude,
abrasive size) have been predicted, assuming that the brittle fracture is pre-
dominant mechanism of metal removal. Experimental verification of the
predicted results from the model has also been obtained in this work.
Liu et al. (2012a) proposed rotary ultrasonic elliptical machining
(RUEM) technique using core drill for cost-effective machining of CFRP
composites. The developed model combined the advantages of core-drill
and elliptical tool vibration towards achieving better quality, delimanation
free holes. The results showed that compared to conventional drilling, the
chip removal rate is improved, tool wear is reduced, precision and surface
quality of drilled hole is improved and sufficient reduction in cutting force
has been achieved.
Wu et al. (2011) presented a stochastic modeling and analysis
technique known as data dependant systems (DDS) to study RUM generated
surface profiles and cutting force signals. Variations in the data sets of sur-
face profiles, for the entrance and exit segments of the machined holes
and cutting force signals were modeled and decomposed to gain insight into
the mechanism of RUM process. It was reported that the roughness of
machined surface is increased as the tool moves deeper due to reduced
flushing efficiency. Moreover, spindle speed and feed rate parameters were
reported to affect the cutting force to a significant extent.
Pei et al. (1995b) have given a review on Rotary Ultrasonic Machining
of structural ceramics. The issues concerning process development, process
modelling and directions of future research for the process have been
undertaken. The material removal mechanism in rotary ultrasonic machin-
ing of ceramics has been detailed out by Pei et al. (1995a). In this article,
experimental evidence for plastic flow as a mode of material removal in
rotary ultrasonic machining has been shown.
Pei and Ferreira (1998) developed two models for prediction of ductile
mode material removal for rotary ultrasonic machining on the assumption
that the brittle fracture is the dominating mode of material removal. The
work material used was magnesia stabilized zirconia for demonstrating
the capability of the developed models for predicting the material removal
for both brittle fracture and ductile failure modes.
Treadwell et al. (2002) presented the Modelling of material removal
rate in rotary ultrasonic machining. An approach to model the MRR during
rotary ultrasonic machining of ceramic has been proposed and applied to
predict the MRR for the case of magnesia stabilized zirconia. In this article,
a five-factor, two-level factorial design was used to study the relationship
Ultrasonic Machining 359

between MRR and the controllable machining parameters. Ya et al. (2002)


have presented an Analysis of the rotary ultrasonic machining mechanism.
The movement of free abrasive particles in the tool tip has been analyzed.
The crack propagation was investigated by using the theory of fracture
mechanics. A mathematical model for the material removal rate has been
proposed to provide a theoretical basis for the analyzing of the process
of rotary ultrasonic machining.
Zeng et al. (2005) investigated the tool wear in rotary ultrasonic
machining of silicon carbide. The effects of input variables on tool wear
and cutting force were assessed through experiments. The tool wear
reported on the end face was much severe than the lateral face. It was also
observed that the diamond grains were dislodged from the tool due to
bond fracture. Figure 15 shows a schematic diagram of RUM process.
The tool wear was found to occur in two stages- first the attritious wear
and then the bond fracture.
Li et al. (2005a) utilized a three variable, two level full factorial design
for investigating the effect of spindle speed, ultrasonic power and feed rate

FIGURE 15 Rotary ultrasonic machining components (Zeng et al., 2005).


360 J. Kumar

on cutting force, MRR and hole quality in RUM of ceramics matrix


composites. The comparison of results with diamond drilling showed
significant reduction in cutting force (50%) over conventional diamond
drilling, while the MRR was improved. All the parameters and their interac-
tions were statistically significant for their effects on hole quality.
Gong et al. (2010) conducted an experimental study on the mechanism
of side milling in rotary ultrasonic machining (RUM). The study showed
that RUM involves lesser amount of tool wear than grinding at the lateral
direction of the cutter under the same conditions. The kinematics of dia-
mond grits was employed for the theoretical analysis. In addition, by using
slim diamond cutters, two different machining strategies were suggested to
side mill a microstructure and grooves on a semi-sphere. An improvement
strategy of material removal rate (MRR) was also devised.
Feng et al. (2012) have reported a feasibility study on rotary ultrasonic
machining of CFRP. Chips, edge chipping, surface roughness, tool wear,
and thrust force were measured under varied conditions of rotation
speed, amplitude of vibration and feed rate. Experimental results showed
that RUM could be used to drill holes in CFRP with high quality and low
tool wear. A better surface was produced by higher rotational speed
and lower feed rate. Cong et al. (2012b) investigated the phenomenon
of edge chipping in RUM of silicon. It was emphasized that holes of dif-
ferent sizes need to be drilled in solar panels made of silicon material and
using conventional methods might lead to cracking of the panel. Two
level three factor full factorial design was employed for planning the
experiments. It was observed that higher tool rotation speed, higher ultra-
sonic power, and lower feed rate led to smaller edge chipping and lower
cutting force. Churi et al. (2009) studied the effects of machining vari-
ables such as spindle speed, feed rate and ultrasonic power on three out-
put variables namely, cutting force, MRR and surface roughness in RUM
of titanium 6Al-4V alloy. Results indicated that spindle speed has signifi-
cant effect on cutting force and surface roughness, but its effects on
MRR were not significant. Feed rate showed significant effects on all
the output variables. The surface roughness was found to decrease while
the ultrasonic power was increased.
Khoo et al. (2008) presented a review on RUM of advanced ceramics.
They reported that the edge chipping is an inevitable phenomenon in drill-
ing hard and brittle materials using RUM but the edge chipping thickness
could be decreased by increasing the support length. The tool wear in
terms of diamond grain dislodgement was concluded to be more predomi-
nant on end face than the lateral face. Wang et al. (2009) investigated the
RUM of potassium dihydrogen phosphate (KDP) crystal with surface
roughness as ouput variable. It was found that the the roughness obtained
by using a tool with chamfered corner was lower than using tools with
Ultrasonic Machining 361

right-angle corners. It was also reported that in the range of


2000–6000 rpm, a higher spindle speed produced higher roughness.
Li et al. (2005b) performed research on the edge-chipping reduction in
rotary ultrasonic machining of ceramics using finite element analysis. Two
failure criterion namely, maximum normal stress criterion and Von Mises
stress criterion were used to predict the relation between the edge chipping
thickness and the support length. It was concluded that as the cutting
depth is increased, the maximum values of normal stress and von Mises
stress increase. Li et al. (2007) reported an experimental study on RUM
of graphite=epoxy panel. They observed that the drilled hole exit edge
quality could be improved significantly with the help of ultrasonic
vibration. The cutting force was found to decrease significantly when ultra-
sonic vibration was introduced. The spindle speed did not affect the aver-
age cutting force and hole entrance=exit edge quality.
Shen et al. (2008) applied support vector Fuzzy adaptive network for
modeling MRR in rotary ultrasonic machining. The fuzzy rules in the
model were generated from the support vectors, which are extracted from
the learning data using SVM algorithm. The combined mechanism connec-
ted SVM and FAN theoretically, improving the performance of newly pro-
posed technique over FAN III in terms of convergent rate, prediction
accuracy. Pei et al. (1995) developed a mechanistic model for prediction
of MRR. A model parameter (which models the ratio of fractured volume
to indented volume of single diamond particle) was shown to be invariant
for most machining conditions. The accuracy of the model was demon-
strated for magnesia stabilized zirconia.
Cong et al. (2011) proposed a novel method for measurement of
vibration amplitude in RUM. It was concluded that excluding cutting tool,
ultrasonic power was the only input variable that significantly affected the
vibration amplitude. Vibration amplitude showed no significant variations
with changes in work material, tool rotational speed and feed rate. Cong
et al. (2012c) presented a study on comparing the machining performance
of RUM, while machining CFRP composites under two different
coolants-cutting fluid and cold air. Cutting force, torque, surface rough-
ness, burning of machined surface, and tool wear were observed under
the conditions of both coolants. The machining conditions under which
cold air would be preferable (for hole quality) to cutting fluid were
recognized.

ULTRASONIC POLISHING
Ultrasonic polishing is a very similar process to ultrasonic machining.
Here, a moving table is employed in addition to a PC-based controller
362 J. Kumar

(Hocheng and Kuo, 2002a). The vibration is generated at higher frequency


and small amplitude, which favors the polishing. Ultrasonic polishing is
very useful for mold applications where skillful handwork is required other-
wise for getting the required degree of finishing. The ultrasonic tool moves
in a patterned path to cover the entire surface to be polished. Microcutting,
plowing and indentation by abrasives could be observed on the polished
surface. Figure 16 shows a typical ultrasonic polishing system.
Hocheng and Kuo (2002b) reported the use of ultrasonic polishing for
mold steel. The experiment was conducted on typical mold steel with the
emphasis on the effects of the abrasive size and the static load on the surface
finish. The optimal abrasive size depending on the surface roughness prior
to polishing was identified to be smaller than the amplitude of vibration and
three to four times the surface roughness of the workpiece. It was also
reported that better finish can be obtained by increasing the static load.
Allen et al. (1995) reported the application of ultrasonic polishing as a
flexible, controllable finishing process for removal of residual stress and
improvement of the surface integrity of the electrodes machined by
EDM. It was concluded that ultrasonic polishing could be useful in improv-
ing the surface properties (such as fatigue strength) of the components
machined by thermal-based processes.
Jones and Hull (1998) presented a combined ultrasonic machining and
abrasive flow machining process, named ‘‘ultrasonic flow polishing’’. They
reported the process to have a good capability for producing a high-quality
finish on the surface of the cavity while causing minimal deterioration to
its profile or dimensional accuracy. The abrasive=polymer mix is pumped
down the centre of the ultrasonically energised tool and on exit, its flow con-
strained by the tool and the workpiece, the mix flows radially relative to the
axis of the tool. While thus constrained, the mix is ultrasonically energized by

FIGURE 16 Ultrasonic polishing machine (Hocheng and Kuo, 2002).


Ultrasonic Machining 363

the vibrating tool. The combination of flow and vibration results in the mix
abrading the workpiece surface. Combination with multi-axis CNC tool
manipulation allows the polishing of complex three-dimensional cavities.

HYBRIDIZATION OF USM WITH OTHER PROCESSES


Babitsky et al. (2004) highlighted ultrasonic assisted turning of aviation
materials through simulation and experimental study. The suggested finite
element model provides numerical comparison between conventional and
ultrasonic turning of inconel 718 alloy in terms of stress=strain rate, cutting
forces and contact conditions at the work-tool interface. Xu and Han
(1999) outlined piezoelectric actuator based active error compensation of
precision machining. Experimental results have shown that the cutting tool
developed is satisfactory in terms of improved roundness and surface
roughness of the machined workpiece.
Sharma et al. (2003) outlined a new longitudinal mode ultrasonic trans-
ducer with an eccentric horn for micro-machining. This device can produce
an angular vibration of the order of 40 kHz at the cutting tip attached at the
end of the horn. The vibrating tip can be used for precision machining of
straight micro-grooves, which are difficult to cut using existing precision
machine tools. The appearance of machined surface has been found to
be excellent when compared with that obtained in conventional grove
machining. Kai and Takahira (1999) highlighted micro-machining by the
application of workpiece vibration. Using this setup, successful machining
of micro-holes in quartz, glass and silicon substrates has been achieved.
Ishikawa et al. (1998) contrived a new drilling method that combines
ultrasonic vibrations of a diamond core drill and low-frequency vibrations
of the workpiece and produced a combined vibration drilling apparatus
for experimentation. The drilling force was observed to decrease by 70%
as compared to conventional drilling using this newly developed method.
Roughness of the drilled hole was found to be higher than non-vibration
drilling, but chipping at the edge of the hole was reduced significantly.
Chang and Bone (2005) outlined the application of ultrasonic vibration
for burr size reduction in the drilling process. High-frequency and
low-amplitude vibration were added in the feed direction during drilling
of the workpiece, which was composed of aluminium.
The results demonstrated that under suitable ultrasonic vibration
conditions, the burr height and width could be reduced in comparison
to conventional drilling. Figure 17 depicts the wavy chip produced under
low frequency condition of the material. Figure 18 shows the variation in
average burr height with regard to the spindle speed under defined experi-
mental conditions.
364 J. Kumar

FIGURE 17 Wavy chip produced in ultrasonic assisted drilling at low frequency (Chang and Bone,
2005).

Azarhoushang and Akbari (2007) presented the design of an


ultrasonically vibrated tool holder and the experimental investigation
of ultrasonically assisted drilling of Inconel 738-LC. The circularity, cylindri-
city, surface roughness and hole oversize of the ultrasonically and conven-
tionally drilled workpieces were measured and compared. The obtained
results showed that the application of ultrasonic vibration can improve
the hole quality considerably.
Liang et al. (2010) proposed a new two-dimensional ultrasonic assisted
grinding (UAG) method and tested its performance in monocrystal silicon
machining. In this method, the workpiece attached on an elliptical ultra-
sonic vibrator was ground with a resin bond diamond grinding wheel under

FIGURE 18 Average burr height vs. spindle speed in ultrasonic assisted drilling (Chang and Bone,
2005).
Ultrasonic Machining 365

the presence of elliptical ultrasonic vibration. The elliptical ultrasonic


vibrator was produced by bonding a piezoelectric ceramic device (PZT)
on a metal elastic body (stainless steel, SUS304) and its detailed
structure=dimensions were determined by finite element method (FEM)
analysis.
Grinding experiments carried out involving monocrystal silicon to con-
firm the performance of the proposed elliptical UAG. In addition, grinding
experiments under the presence of the axial and vertical ultrasonic vibra-
tions and the absence of ultrasonic vibrations, i.e., conventional grinding,
were also carried out for comparison. The obtained results showed that
compared with conventional grinding, the axial ultrasonic vibration results
in greatest improvement in the work surface quality and a slight reduction
in the grinding forces. The elliptical ultrasonic vibration leads to the signifi-
cant reduction of both the surface roughness and grinding forces.
Tabatabaei et al. (2013) presented an analysis of ultrasonic assisted
machining (UAM) on regenerative chatter in turning. Time domain
numerical analysis of chatter under ultrasonic vibration was developed by
integrating the equations of orthogonal cutting in lathe with ultrasonic
assisted cutting equations. The obtained results from the numerical analysis
for some selected points of the stability lobe were compared and validated
against the observed experimental results. It was shown that ultrasonic
vibration can improve the stability for some cutting conditions, while
degrading the stability in some other conditions.
The application of USM assisted turning for titanium and its alloys, a
ductile and tough material, has been reported by few researchers. Sharman
et al. (2001) outlined the application of ultrasonic assisted turning to
c-titanium aluminide. As compared to conventional turning, the cutting
forces were reported to be of very small magnitude (approx. 12%) in this
process, thereby, improving both the tool life and surface finish. Aspinwall
and Kasuga (2001) have reported the use of USM for production of 3 mm
holes in titanium aluminide.
Although machining with conventional methods, titanium aluminide;
an alloy of titanium, encounters problems of surface integrity and
micro-cracking. When machined with USM, satisfactory results have been
achieved with polycrystalline diamond tooling. Grit size has been identified
as the greatest factor affecting MRR followed by static load; tool type: solid=
hollow and power level. In contrast to brittle materials, the combination of
fine grit size, low power level and solid tool type gives maximum TWR.
Mohsen et al. (2012) reported a novel technique for ultrasonic assisted
grinding of Ti-6Al-4V alloy. In this research, the effect of imposition of ultra-
sonic vibration on the grinding of Ti6Al4V alloy was studied. Longitudinal
vibration at ultrasonic frequency range (20 kHz) was applied on the work-
piece and machining forces and surface roughness were compared between
366 J. Kumar

conventional grinding (CG) and ultrasonic assisted grinding (UAG)


processes. An ultrasonic setup was designed, optimized and fabricated based
on combination of mathematical modeling, FEM analysis and genetic algor-
ithm. Comparison between CG and UAG at several cutting and feed speeds
and cutting depths showed reduction of grinding forces and improvement
of surface roughness as a result of ultrasonic application.
Choi, Jeon, and Kim (2007) explored the use of chemical-assisted ultra-
sonic machining for machining of glass. To obtain the chemical effects, a
low concentration hydrofluoric acid solution was added to the slurry. The
hybrid method was found to be superior in terms of material removal rate
and integrity of the machined surface. The surface roughness was also
found to improve and the machining load was decreased significantly as
compared to the conventional USM method. Lau et al. (1994) investigated
the ultrasonic-aided laser drilling of alumina-based metal matrix
composites.
The conventional pulsed Nd: YAG laser drilling involved problems of
limited depth of drill, non-cylindrical hole profile and the presence of
excessive recast layer. The hybrid ultrasonic-laser drilling technique was
found to improve the capabilities in deep hole drilling with more precision.
The thickness of heat affected zone (HAZ) was reduced by 30% under the
action of ultrasonic vibration which aided in effective utilization of the laser
energy in vaporizing the material. Tsutsumi, Okano, and Suto (1993) inves-
tigated the machining performance of ultrasonic-aided core drilling
method for machining of ceramics. The thrust force was reported to
decrease by 70% without affecting the surface quality of the machined hole.
This was contributed to the more effective penetration of diamond grits
into the surface of work material, reducing the required drilling thrust
force. The width of the chip produced was increased and the chipping of
the edges was found to be lowered.
Many investigators have explored hybrid EDM-USM process for
machining of tough materials and advanced ceramics (Lin et al., 2000;
Wansheng et al., 2002; Yan and Biing, 2001; Zhinxin et al., 1997). The
results from these investigations reveal a superior performance of the
hybrid method in terms of improved machining rate, reduced heat
affected zone and recast layer along with better stability and control
when compared to the conventional EDM process. This has been attrib-
uted to effective flushing of the debris and wear particles from the
machining zone due to the action of ultrasonic vibration, which
improves process stability and promotes more efficient heat transfer
from the work surface.
Wansheng et al. (2002) investigated the effect of ultrasonic vibration
introduction in EDM process to machine micro-holes in titanium alloy
Ti-6Al-4V; concluding an increase in MRR as well as the process stability
Ultrasonic Machining 367

along with reduction in arcing phenomenon. When applied to machining


of titanium alloy, the combined EDM-USM process has been found to dem-
onstrate better performance in terms of improved MRR, discharging
efficiency and reduced thickness of the recast layer (Wansheng et al.,
2002). Lin et al. (2000) have also shown a similar result; the introduction
of ultrasonic vibration in EDM for deep hole drilling in titanium alloy
(Ti-6Al-4V) can improve the machining quality and efficiency distinctly.
This is due to the fact that the debris and small particles generated by ero-
sion of the work surface are efficiently disposed by the ultrasonic vibration
induced, which gives a better stability to the process. Zhinxin et al. (1997)
studied a combined EDM-USM machining of titanium alloy. Results show
that the machining efficiency obtained was 2–3 times higher than conven-
tional EDM process. The surface integrity of the machined samples was also
found to be better when compared to conventional EDM.
Chen and Lin (2009) investigated the surface modification of Al-Mg-Zn
alloy using combined EDM-USM process while adding TiC particles in the
dielectric. The elemental distributions of titanium and carbon on the
cross-section were quantitatively determined using an electron probe
micro-analyzer (EPMA). Micro-hardness and wear resistance tests were con-
ducted to evaluate the modifications on the machined surface caused by
the combined process. The experimental results showed that the combined
process was associated with improved machining performance. The combi-
nation of EDM with USM yielded an alloyed layer that improved the hard-
ness and wear resistance of the machined surface.
Kang et al. (2012) studied the effect of ultrasonic vibration in nanose-
cond laser machining. The morphological change of particles re-deposited
on the machined area was observed under varied process conditions. The
results showed that the surface finish was improved by the near-field surface
cooling enhancement induced by ultrasonic vibration. From the chemical
composition analysis of the machined surface, it was verified that the
vibration could prevent the surface oxidation and the formation of recast
layer. As a result, the ultrasonic vibration in ns-laser machining process
could be effective for improving both the physical and chemical character-
istics of the machined surface.
Abdullah et al. (2012) investigated the enhancement of strength for
welded structures by applying ultrasonic peening method.
For investigating the effect of ultrasonic peening on stainless steel-304
welded parts, a series of experiments were designed and implemented. This
research comprised the results of experimental fatigue strength tests
along with metallography, micro-hardness and corrosion resistance tests
of welded pieces with and without processing by ultrasonic peening.
Experiments proved that under post treatment by ultrasonic peening, a
better mechanical and corrosion resistance is achieved.
368 J. Kumar

GAPS IDENTFIED FROM LITERATURE REVIEW AND


DIRECTIONS FOR FUTURE RESEARCH

1. The correlation between material removal mode (such as


micro-chipping) and the hole integrity (entrance chipping, wall rough-
ness and sub-surface damage) is needed to be evaluated, for most of
the materials that are reported to be machined by application of static
or rotary USM. This aspect has been investigated by only one or two
researchers, for very limited number of materials (brittle materials)
as the available literature suggests. There is a critical need of perform-
ing experimental studies on ductile and tough materials in this regard,
for which the hole integrity is affected to a larger extent than brittle
materials.
2. The phenomenon responsible for variation of the roughness of drilled
hole along the depth of cut has not been clearly recognized. An in
depth investigation is needed to understand the mechanics of the dif-
ference in roughness at entry and exit parts of the drilled hole, parti-
cularly in rotary mode of USM.
3. There is a critical need for dynamic process modeling of output vari-
ables (MRR, cutting force, etc.) for machining newly developed materi-
als such as advanced composites, MMC’s and high strength to weight
ratio (aerospace) materials with USM. The available literature mainly
focuses on static parametric relationships. Materials such as titanium,
nickel alloys and stellite may be used as work material in this type of
research activity.
4. Non-traditional optimization algorithms such as genetic algorithm,
artificial bee colony (ABC), harmony search (HS) and particle swam
optimization (PSO) can be employed for performing multi-objective
optimization of correlated output variables. The optimized results from
these methods can be compared with other traditional optimization
methods such as desirability function, utility concept, Principal compo-
nent analysis, grey relational analysis.
5. Application of simulation tools and fuzzy logic modeling might be
applied for checking the validity of proposed models as well as for con-
structing the models for reference purpose. The machine operators
and process planners may refer the results obtained by applying these
tools from time to time and hence the time and other resources could
be saved.
6. Further research is needed on obtaining a robust process performance
while machining advanced materials by applying robust design con-
cepts such as Taguchi’s loss function. Most of the studies reported on
parametric optimization of USM have made use of Taguchi’s design
of experiments approach; with a focus on improving the mean value
Ultrasonic Machining 369

of quality characteristics. There is almost no study reported on bring-


ing the process performance on target or controlling the variability
of the process.
7. The interactions among the controllable input factors and uncon-
trollable noise factors such as work material composition, slurry
concentration variation, abrasive wear, tool geometry degradation
may be assessed by systematic variation of the both type of fac-
tors. This would again help in achieving improved process perfor-
mance through quantification of the loss resulting from such
interactions.
8. Process capability studies are needed to be undertaken for enhanc-
ing the process capability index for quality characteristics such as
surface quality. The parametric optimization of process capability
index may be obtained while machining materials of different mech-
anical properties using USM. The process capability study may
include power consumption of the USM setup and its components,
surface quality obtained, limiting value of depth of cut and
reduction of machining efficiency with the progress of machining
(time of machining).
9. The issues of micro-hardness of the machined surface, sub-surface
hardness profile, sub-surface damage (change in fatigue strength,
microstructure) need to be explored for a better assessment of surface
integrity of ultrasonically machined surfaces. These issues are vital for
work materials such as titanium which is used in typical applications
demanding least possible changes in the surface integrity post
machining.
10. Use of USM for drilling a large number of small diameter holes
(perforation) needs to be investigated properly. Machining procedures
and efficient tooling (such as multi-thin tools) regarding this need to
be developed. Improvement of the machining efficiency of USM in this
regard would bring this process close to the level of performance of
other processes such as EBM or LBM.
11. Few parameters such as aspect ratio (ratio of depth of cut to diameter)
of the drilled hole, roughness of the tool face have not been included
in the experimental investigation as input process parameters. Their
effects on process characteristics such as surface quality, geometrical
inaccuracy (out of roundness, conicity) are needed to be assessed for
improving the process capability. Moreover, efficient methods for slurry
flow into the machining zone need to be developed while drilling
deep, fine sized holes using USM as the maximum drilling depth is
usually limited.
12. The cavitation phenomenon and its contribution in the material
removal are needed to be investigated for static mode USM. The effect
370 J. Kumar

of cavitation on surface quality may also be evaluated. No model has


been presented in this regard in the reported literature.
13. Although USM is recognized as a process in which the work material
doesn’t undergo any significant thermal effect, however, there is a
measurable heat input to the tool while vibration is applied to it, parti-
cularly if there is some micro-void inside the tool or at some joint. For
tools composed to low melting point metals, there may be a possibility
of alteration in the mechanical properties or phase change, which is
needed to be investigated.
14. Further research may be focused on improvement of slurry properties
or abrasive action of the slurry by addition of suitable chemical sub-
stances, varying the slurry temperature, by suspending the abrasive in
different types of media.
15. An analysis into the economies of machining with USM is needed to be
performed, particularly from the point of view of tool fabrication cost
and productive life of the tool. Cost-effective methods need to be estab-
lished for tool fabrication.

Figure 19 depicts a summary of the identified potential research areas


for USM.

FIGURE 19 Potential research areas for USM. (Figure available in color online.)
Ultrasonic Machining 371

CONCLUSIONS
Ultrasonic machining is one of the most widely used non-traditional
processes; especially for commercial machining of hard, brittle and
fragile materials. There is tremendous scope for application of USM
for establishing cost effective machining solutions for relatively tough
and ductile metals such as titanium, nickel-alloys. Productivity and qual-
ity of USM process is dependant on the work material properties (such
as hardness and fracture toughness), tool properties (hardness, impact
strength and finish), abrasive properties (hardness, coarseness and
viscosity) and process settings (power input, static load, amplitude
and frequency of vibration). The material removal in USM has been
found to occur by propagation and intersection of median and lateral
cracks that are induced due to repeated impacts of abrasive grains.
In addition, following conclusions could be drawn regarding various
aspects of USM:

1. The material removal follows a mixed mode (brittle fracture of the


work surface preceded by plastic deformation) while a tough material
such as titanium is machined with USM. By manipulating the input
parameter settings, different desirable ratios of the two modes can be
obtained experimentally.
2. Because of strain hardening of the work surface (while machining
ductile materials) during machining, the micro-hardness of the work
surface is increased. Micro-hardness gain is found to be higher when
the process settings that result in lower energy input rate to work
surface are used for machining.
3. The tool surface undergoes a change in shape during machining due
to uneven distribution of abrasive grains under the tool face leading to
non-uniform wear. The change in shape has been found to be less for
very hard tools such as carbide and high speed steel. It is necessary to
restore the tool shape before machining further.
4. Design of the tool is a very crucial factor affecting the productivity of
USM process. Tools with higher mass tend to suppress the amplitude
of vibration thereby reducing the machining rate. Tools with excessive
tip length tend to experience deformation and micro-cracking during
machining. Fatigue failure of the tool takes place if there is a slight
misalignment between the tool and the horn.
5. The tightening of screw attachment with tool horn should be opti-
mum, higher tightening results in to permanent ultrasonic welding
of screw with horn. Proper sized acoustic washer generally made up
of copper or white metal should be used and replaced after every
dismantling of tool=horn assembly for optimum MRR=TWR.
372 J. Kumar

6. There is tremendous scope for further research in many crucial aspects


of USM process that either have not been addressed at all, or the inves-
tigation regarding them lacks a focussed approach. A few of these are:
i. The phenomenon of cavitation and its effects on surface quality and
material removal need to be assessed through dynamic modeling.
ii. Assessment of the momentum imparted to the impinging abrasive
particles under the varied conditions of operating parameters and
evaluation of the subsequent effect on machining performance of
USM.
iii. Further investigation is needed on improving the accessibility of the
slurry in the center of the tool face so that geometrical inaccuracy
in the bottom flat of the drilled holes (convex surface) could be
tackled.
iv. More research effort is needed to be put on exploring the issues
such as hole integrity, delamination of drilled holes, sub-surface
damage, micro-hardness of machined surface, improvement in
the limiting value of aspect ratio of the holes, enhancing the
process capabilities such as surface quality, precision of machining,
power consumption, etc.
v. Development and validation of dynamic models for prediction of
MRR, cutting force and tool life for a wide spectrum of material
properties, tool properties and slurry characteristics is needed.
Simulation tools may be applied for this purpose.

REFERENCES
Abdullah, A.; Massoud, M.; Ahmad, E. (2012) Strength enhancement of welded structures by ultrasonic
peening. Materials and Design, 38: 7–18.
Adithan, M. (1981) Tool wear characteristics in ultrasonic drilling. Tribology International, 14(6):
351–356.
Adithan, M.; Krishnamurthy, R. (1978) Structural alterations in the workpiece by ultrasonic drilling.
Wear, 46: 327–334.
———. (1976) Production accuracy of holes in ultrasonic drilling. Wear, 40(3): 309–318.
———. (1978) An appraisal of wear mechanisms in ultrasonic drilling. Annals of the CIRP, 27: 119–124.
Adithan, M.; Venkatesh, V.C. (1974) Parametric influence on tool wear in ultrasonic drilling. Tribology
International, 7(6): 260–264.
Allen, B.J.; Williams, R.E.; Gilmore, R. (1995) Surface integrity improvement of EDM components by
ultrasonic polishing. Technical Papers of NAMRI=SME: 61–66.
Amin, A.K.M.; Ismail, F.A.; Khairushima, M.K. (2007) Effectiveness of uncoated WC-Co and PCD inserts
in end milling of titanium alloy-Ti-6Al-4V. Journal of Materials Processing Technology, 192–193: 147–
158.
Amin, S.G.; Ahmed, M.H.; Youssef, H.A. (1993) Optimum design charts of acoustic horns for ultrasonic
machining. Proc. Inter. Conference on Advances in Materials Processing Technologies, 24–27 August,
Dublin City University, Dublin, 139–147.
Anantha Ramu, B.L.; Krishnamurthy, R. (1989) Machining performance of toughened zirconia ceramic
and cold compact alumina ceramic in ultrasonic drilling. Journal of Mechanical Working Technology,
120: 365–375.
Ultrasonic Machining 373

Aspinwall, D.K.; Kasuga, V. (2001) The use of ultrasonic machining for the production of holes in Y-TIAL.
Proceedings of the 13th International Symposium for Electro Machining ISEM XIII, Spain, 2: 925–937.
Azarhoushang, B.; Akbari, J. (2007) Ultrasonic assisted drilling of Inconel 738-LC. International Journal of
Machine Tools and Manufacture, 47: 1027–1033.
Babitsky, V.I.; Mitrofanov, A.V.; Silverschmidt, V.V. (2004) Ultrasonically assisted turning of aviation
materials: simulations and experimental study. Ultrasonics, 42: 81–86.
Baek, D.K.; Ko, J.T.; Seung, H.Y. (2013) Enhancement of surface quality in ultrasonic machining of glass
using a sacrificing coating. Journal of Materials Processing Technology, 213: 553–569.
Balamuth, L. (1964) Ultrasonic vibrations assisted cutting tools. Metalworking Production, 108(24): 75–77.
Balamuth, L.A. (1966) Ultrasonic assistance to conventional metal removal. Ultrasonics, 4: 125–130.
Barash, M.M.; Watanapongse, D. (1970) On the effect of ambient pressure on the rate of material
removal in ultrasonic machining. International Journal of Mechanical Sciences, 12: 775–779.
Benedict, G.F. (1987) Non-Traditional Manufacturing Processes, Marcel Dekker Inc., New York, pp. 67–86.
Bhattacharya, A. (1973) New Technology, The Institution of Engineers (I), Calcutta, pp. 12–17.
Bulat, T.J. (1974) Micro-Sonics in Industry: Ultrasonic Cleaning, Bendix and Life Supports Division
Publication, USA, 120.10.153: 13.
Chang, S.; Bone, G.M. (2005) Burr size reduction in drilling by ultrasonic assistance. Robotics and
Computer-Integrated Manufacturing, 120: 442–450.
Chen, Y.-F.; Lin, Y.-C. (2009) Surface modifications of Al-Zn-Mg alloy using combined EDM with ultra-
sonic machining and addition of TiC particles into the dielectric. Journal of Materials Processing
Technology, 209: 4343–4350.
Choi, J.P.; Jeon, B.H.; Kim, B.H. (2007) Chemical-assisted ultrasonic machining of glass. Journal of
Materials Processing Technology, 191: 153–156.
Churi, N.J.; Pei, Z.J.; Treadwell, C. (2009) Rotary ultrasonic machining of titanium alloy. Machining
Science and Technology, 10(3): 301–321.
Cong, W.L.; Pei, A.J.; Deines, T.W.; Srivastava, A.; Riley, L.; Treadwell, C. (2012a) Rotary ultra-
sonic machining of CFRP composites: A study on power consumption. Ultrasonics, 52(8):
1030–1037.
Cong, W.L.; Pei, Z.J.; Feng, Q.; Deines, T.W.; Treadwell, C. (2012b) Edge chipping in rotary ultrasonic
machining of silicon. International Journal of Manufacturing Research, 7(3): 311–329.
———. (2012c) Rotary ultrasonic machining of carbon reinforced fiber plastic composites: using cut-
ting fluid versus cold air as coolant. Journal of Composite Materials, 46(14): 1745–1753.
Cong, W.L.; Pei, Z.J.; Mohanty, N.; Van Vleet, E.; Treadwell, C. (2011) Vibration Amplitude in Rotary
Ultrasonic Machining: A Novel measurement method and effects of process variables. Journal of
Manufacturing Science and Engineering, 133: 034501–1–034501–6.
Cook, N.H. (1966) Manufacturing Analysis, Addison-Wesley, New York, pp. 133–148.
Dam, H.; Jensen, J. (1993) Surface characterization of ultrasonic machined ceramics with diamond
impregnated sonotrode. Machining of Advanced Materials, 34: 125–133.
Dam, H.; Quist, P.; Schreiber, M. (1995) Productivity, surface quality and tolerances in ultrasonic
machining of ceramics. Journal of Materials Processing Technology, 51(1–4): 358–368.
Deng, J.; Lee, T. (2000) Surface integrity in electro-discharge machining, ultrasonic machining and
diamond saw cutting of ceramic composites. Ceramic International, 26(8): 825–830.
Deng, J.; Lee, T. (2002) Ultrasonic machining of alumina based ceramic composites. Journal of the
European Ceramic Society, 22(8): 1235–1241.
Dvivedi, A.; Kumar, P. (2007) Surface quality evaluation in ultrasonic drilling through the Taguchi
technique. International Journal of Advanced Manufacturing Technology, 34(1–2): 131–140.
Farago, F.T. (1980) Abrasives methods engineering. Industrial Press, 2: 480–481.
Farzin-Nia, F.; Sterrett, T. (1990) Effect of machining on fracture toughness of corundum. Journal of
Materials Science, 25(5): 2527–2531.
Feng, Q.; Cong, W.L.; Pei, Z.J.; Ren, C.Z. (2012) Rotary ultrasonic machining of carbon fiber reinforced
polymer: feasibility study. Machining Science and Technology, 16(3): 380–398.
Frederick, J.R. (1965) Ultrasonic Engineering, John Wiley and Sons Inc., New York, pp. 32–45.
Gauri, S.K.; Chakravorty, R.; Chakraborty, S. (2011) Optimization of correlated multiple responses of
ultrasonic machining (USM) process. International Journal of Advanced Manufacturing Technology,
53: 1115–1127.
374 J. Kumar

Ghabrial, S.R. (1986) Trends towards improving the surfaces produced by modern processes. Wear,
109(1–4): 113–118.
Ghabrial, S.R.; Saleh, S.M.; Kohail, A.; Moison, A. (1982) Problems associated with electro-discharge
machined, electro-chemically machined and ultrasonically machined surfaces. Wear, 83: 275–283.
Ghahramani, B.; Wang, Z.Y. (2001) Precision ultrasonic machining process: A case study of stress analy-
sis of ceramic (Al2O3). International Journal of Machine Tools and Manufacture, 41(8): 1189–1208.
Gilmore, R. (1991) Ultrasonic machining—a case study. Journal of Materials Processing Technology, 28(1–2):
139–148.
Gilmore, R. (1989) Ultrasonic machining and orbital abrasion techniques. SME Technical Paper (series)
AIR, NM89–419: 1–20.
Goetze, D. (1956) Effect of vibration amplitude, frequency, and composition of the abrasive slurry on
the rate of ultrasonic machining in Ketos Tool Steel. Journal of Acoustical Society of America, 28(6):
1033–1045.
Gong, H.; Feng, F.Z., Hu, X.T. (2010) Kinematic view of tool life in rotary ultrasonic side milling of hard
and brittle materials. International Journal of Machine Tools and Manufacture, 50(3): 303–307.
Graff, K.F. (1975) Macrosonics in industry: Ultrasonic machining. Ultrasonics, 13: 103–109.
Guzzo, P.L.; Raslan, A.A.; De Mello, J.D.B. (2003) Ultrasonic abrasion of quartz crystals. Wear, 255:
67–77.
Halm, R.; Schulz, P. (1993) Ultrasonic machining of complex ceramic components. Erosion AC Report,
DKG 70(7): 6.
Haun, R.; Schulz, P. (1994) New ultrasonic machining route to create complex ceramic components.
Proc. IEEE Ultrasonic Symposium, 31 Oct.–3 Nov., Cannes, France, 3: 1389–1392.
Hocheng, H.; Kuo, K.L. (2002a) On-line tool wear monitoring during ultrasonic machining using tool
resonance frequency. Journal of Materials Processing Technology, 123(1): 80–84.
Hocheng, H.; Kuo, K.L. (2002b) Fundamental study of ultrasonic machining of mold steel. International
Journal of Machine Tools and Manufacture, 42: 7–13.
Hocheng, H.; Kuo, K.L.; Lin, J.T. (1999) Machinability of zirconia ceramics in ultrasonic drilling.
Materials and Manufacturing Processes, 14(5): 713–724.
Hocheng, H.; Hsu, C.C. (1995) Preliminary study of ultrasonic drilling of Fiber-Reinforced Plastics. Jour-
nal of Materials Processing Technology, 48: 255–266.
Hu, P., Zhang, J.M., Pei, Z.J., Treadwell, C. (2002) Modeling of material removal rate in rotary ultrasonic
machining: designed experiments. Journal of Materials Processing Technology, 129: 339–344.
Ishikawa, K.; Suwabe, H.; Nishide, T.; Uneda.M. (1998) A study on combined vibration drilling by ultra-
sonic and low-frequency vibrations for hard and brittle materials. Precision Engineering, 22: 197–206.
Jadoun, R.S.; Kumar, P.; Mishra, B.K.; Mehta, R.C.S. (2006) Manufacturing process optimization for tool
wear rate in ultrasonic drilling of engineering ceramics using the Taguchi method. International
Journal of Machining and Machinability of Materials, 1(3–4): 94–114.
Jain, N.K.; Jain, V.K. (2001) Modeling of material removal in mechanical type of advanced machining
processes—a state of the art review. International Journal of Machine Tools and Manufacture, 41:
1573–1635.
Jones, A.R.; Hull, J.B. (1998) Ultrasonic flow polishing. Ultrasonics, 36: 97–101.
Kaczmarek, I. (1976) Principles of Machining by Cutting, Abrasion and Erosion, Peter Peregrinus Ltd.,
Stevenage, pp. 448–462.
Kai, E.; Takahira, M. (1999) Micro ultrasonic machining by application of work-piece vibration. CIRP
Annals, 48(1): 131–134.
Kainth, G.S.; Nandy, A.; Singh, K. (1979) On the mechanisms of material removal in ultrasonic machin-
ing. International Journal of Machine Tool Design, 19: 33–41.
Kang, B.; Kim, W.G.; Yang, M.; Cho, S.-H.; Park, J.-K. (2012) A Study of the effect of ultrasonic vibration
in nanosecond laser machining. Optics and Lasers in Engineering, 50: 1817–1822.
Kazantsev, V.F. (1966) Improving the output and accuracy of ultrasonic machining. Machines and Tooling,
37(4): 33–39.
Kazantsev, V.F. (1963) The relationship between output and machining conditions in ultrasonic machin-
ing. Machines and Tooling, 34: 14–17.
Kennedy, D.C.; Grieve R.J. (1975) Ultrasonic machining—A review. The Production Engineer, 54(9):
481–486.
Ultrasonic Machining 375

Kennedy, W.J.; Sakaar, C. (1989) Improving the machining of ceramics. SME Technical Paper, Orlando,
FL, USA, 6–9..
Khairy, A.B.E. (1990) Assessment of some dynamic parameters for the ultrasonic machining process.
Wear, 137(2): 187–198.
Khoo, C.Y.; Hamzah, E.; Sudin, I. (2008) A review on the rotary ultrasonic machining of advanced
ceramics. Journal Mekenikal, 25: 9–23.
Kohals J.B. (1984) Ultrasonic manufacturing process-ultrasonic machining and ultrasonic impact grind-
ing (USIG). The Carbide and Tool Journal, 16(5): 12–15.
Komaraiah, M.; Reddy, P.N. (1993a) A study on the influence of workpiece properties in ultrasonic
machining. International Journal of Machine Tools and Manufacture, 33: 495–505.
Komariah, M.; Reddy, P.N. (1993b) Relative performance of tool materials in ultrasonic machining.
Wear, 161(1–2): 1–10.
Komaraiah, M.; Manan, M.A.; Reddy, P.N.; Victor, S. (1988) Investigation of surface roughness and accu-
racy in ultrasonic machining. Precision Engineering, 10(2): 59–68.
Koops, L. (1964) Investigation into the influence of the wear of abrasive powder on the technological
indices of ultrasonic machining. Annals of the CIRP, 13(3): 151–157.
Koval Chenko, M.S.; Paustovskii, A.V.; Perevyazko, V.A. (1986) Influence of properties of abrasive mate-
rials on the effectiveness of ultrasonic machining of ceramics. Powder Metallurgy and Metal Ceramics,
25: 560–562.
Kremer, D. (1981) The state of the art of ultrasonic machining. Annals of the CIRP, 30: 107–115.
Kremer, D.; Bazine, G.; Moison, A. (1983) Ultrasonic machining improves EDM technology, Electro
machining. Proc. 7th Int. Symp., Birmingham, UK, 67–76.
Kremer, D.; Mackie, J. (1988) Ultrasonic machining applied to ceramic materials. Industrial Ceramics,
8(3): 632–637.
Kremer, D. (1991) New developments in ultrasonic machining. SME Technical Paper, MR91–522: 13.
Kubota, M.; Tamura, Y.; Shimamura, N. (1977) Ultrasonic machining with a diamond impregnated tool.
Bulletin of Japanese Society of Precision Engineering, 11(3): 127–132.
Kumar, J.; Khamba, J.S. (2008) An experimental study on ultrasonic machining of pure titanium using
designed experiments. Journal of the Brazilian Society of Mechanical Sciences and Engineering, 30(3):
231–238.
Kumar, J.; Khamba, J.S.; Mohapatra, S.K. (2008) An investigation into the machining characteristics of
titanium using ultrasonic machining. International Journal of Machining and Machinabaility of
Materials, 3(1–2): 143–161.
Kumar, J.; Khamba, J.S.; Mohapatra, S.K. (2009) Investigating and modeling tool-wear rate in the ultra-
sonic machining of titanium. International Journal of Advanced Manufacturing Technology, 41(11):
1101–1111.
Kumar, J.; Khamba, J.S. (2009) An Investigation into the effect of work material properties, tool
geometry and abrasive properties on performance indices of ultrasonic machining. International
Journal of Machining and Machinability of Materials, 5(2–3): 347–365.
Kumar, J.; Khamba, J.S. (2010a) Modeling the material removal rate in ultrasonic machining of titanium
using dimensional analysis. International Journal of Advanced Manufacturing Technology, 48(1–4): 103–119.
Kumar, J.; Khamba, J.S. (2010b) Multi-Response optimization in ultrasonic machining of titanium
using Taguchi’s approach and utility concept. International Journal of Manufacturing Research,
5(2): 1013–1022.
Kumar, A.; Kumar, V.; Kumar, J. (2012) Prediction of surface roughness in wire electric discharge
machining (WEDM) process based on response surface methodology. International Journal of
Engineering and Technology, 2(4): 708–712.
Kumar, A.; Kumar, V.; Kumar, J. (2013a) Investigation of machining parameters and surface integrity in
wire electric discharge machining (WEDM) of pure titanium. InProceedings of the Institution of Mech-
anical Engineers, Part B: Journal of Engineering Manufacture, DOI:10.1177/0954405413479791.
——— (2013b) Multi-Response optimization of process parameters based on response surface method-
ology for pure titanium using WEDM process. InInternational Journal of Advanced Manufacturing
Technology, DOI:10.1007/s00170-013-4861-9.
Kumehara, H. (1984) Characteristics of threaded joints in ultrasonic vibrating system. Bulletin of JSME,
27(223): 117–123.
376 J. Kumar

Lalchnuanvela, H.; Doloi, B.; Bhattacharya, B. (2012) Enabling and understanding ultrasonic machin-
ing of engineering ceramics using parametric analysis. Materials and Manufacturing Processes, 27(4):
443–448.
Laroiya, S.C.; Adithan, M. (1993) Tool wear in machining of advanced ceramics. Xth National Conference
on Industrial Tribology, March 1993, conducted at Indian Institute of Petroleum, Dehradun, India,
164–169.
Lau, W.S.; Yue, T.M.; Wang, M. (1994) Ultrasonic-aided laser drilling of aluminium based metal matrix
composites. Annals of the CIRP, 43: 177–182.
Lawn, B.R.; Evans, A.G.; Marshall, D.B. (1980) Elastic=plastic indentation damage in ceramics: the med-
ian=radial crack system. Journal of the American Ceramic Society, 63(9–10): 574–581.
Lee, T.C.; Chan, C.W. (1997) Mechanism of the ultrasonic machining of ceramic composites. Journal of
Materials Processing Technology, 71: 195–201.
Li, Z.C.; Jiaoa, Y.; Deinesa, T.W.; Pei, Z.J. (2005a) Rotary ultrasonic machining of ceramics matrix com-
posites: feasibility study and designed experiments. International Journal of Machine Tools and Manu-
facture, 45(12–13): 1402–1411.
Li, Z.C.; Cai, Wu-L.; Pei, Z.J.; Treadwell, C. (2005b) Edge chipping reduction in rotary ultrasonic
machining of ceramics: finite element analysis and experimental verification. International Journal
of Machine Tools and Manufacture, 46(12–13): 1469–1477.
Li, Z.C.; Sisco, T.; Pei, Z.J.; Micale, A.C.; Treadwell, C. (2007) Experimental study on rotary ultrasonic
machining of graphite=epoxy panel. Proceedings of the ASPE 2007 Spring Topical Meeting on Vibration
Assisted Machining Technology, Chapel Hill, NC, April 16–17: 52–57.
Liang, Z.; Wu, Y.; Wang, X.; Zhao, W. (2010) A new two dimensional ultrasonic assisted grinding (UAG)
method and its fundamental performance in monocrystal silicon machining. International Journal of
Machine Tools and Manufacture, 50: 728–736.
Lin, Y.C.; Yan, B.H.; Chang, Y.S. (2000) Machining characteristics of titanium alloy (Ti-6Al-4V) using a
combination process of EDM with USM. Journal of Materials Processing Technology, 104: 171–177.
Liu, J.; Zhang, D.; Qin, L.; Yan, L. (2012a) Feasibility study of the rotary ultrasonic elliptical machining
of carbon fiber reinforced plastics (CFRP). International Journal of Machine Tools & Manufacture, 53:
141–150.
Liu, D.; Cong, W.L.; Pei, Z.J.; Tang, Y. (2012b) A cutting force model for rotary ultrasonic machining of
brittle materials. International Journal of Machine Tools & Manufacture, 52: 77–84.
Majeed, A.A.; Vijayaraghvan, L.; Malhotra, S.K.; KrishnaMurthy, R. (2008) A.E. monitoring of ultrasonic
machining of Al2O3=LaPO4 composites. Journal of Materials Processing Technology, 207: 321–329.
Majeed, M.A.; Vijayaraghvan, L.; Malhotra, S.K.; KrishnaMurthy, R. (2008) Ultrasonic machining of
Al2O3=LaPO4 composites. International Journal of Machine Tools & Manufacture, 48: 40–46.
Markov, A.I. (1959) Kinematics of the dimensional ultrasonic machining method. Machines and Tooling,
30(10): 28–31.
Markov, A.I. (1977) Ultrasonic drilling and milling of hard non-metallic materials with diamond tools.
Machines and Tooling, 48(9): 45–47.
Marshall, D.B.; Lawn, B.R.; Evans, A.G. (1982) Elastic=plastic indentation damage in ceramics: the
lateral crack system. Journal of the American Ceramic Society, 65(11): 561.
McGeough, J.A. (1988) Advanced Methods of Machining, Chapman and Hall, London, ISBN 0412319705:
pp. 170–198.
Merkulov, L.G. (1957) Design of ultrasonic concentrations. Akusticheskiy Zhurnal, 3: 246–255.
Miller, G.E. (1957) Special theory of ultrasonic machining. Journal of Applied Physics, 28(2): 149–156.
Mishra, P.K. (2005) Non-Conventional Machining, Narosa Publishing House, New Delhi, pp. 22–44.
Mohsen, N.G.; Mohammed, M.G.; Javed, A. (2012) Ultrasonic Assisted Grinding of Ti6Al4V alloy.
Procedia CIRP, 1: 353–358.
Moore, D. (1985) Ultrasonic impact grinding. Proc. Non-Traditional Machining Conference, Cinicinnati,
OH, USA, 137–139.
Moreland, M.A. (1987) Ultrasonic Advantages revealed in ‘‘the hole story.’’ Creative Manufacturing
Engineering Program, Society of Manufacturing Engineers, pp. 1–7.
Moreland, M.A. (1988b) Versatile performance of ultrasonic machining. Ceramic Bulletin, 67(6): 1045–1047.
Nair, E.V.; Ghosh, A. (1985) A fundamental approach to the study of mechanics of ultrasonic
machining. International Journal of Production Research, 23: 731–753.
Ultrasonic Machining 377

Nath, C.; Lim, G.C.; Zheng, H.Y. (2012) Influence of the material removal mechanisms on hole integrity
in ultrasonic machining of structural ceramics. Ultrasonics, 52: 605–613.
———. (1964) Ultrasonic machining and forming. Ultrasonics, 2: 167–173.
———. (1972) Macrosonics in industry. Ultrasonics, 10: 9–13.
Neppiras, E.A. (1957) Ultrasonic Machining-II. Operating conditions and performance of ultrasonic
drills. Philips Technology Review, 18(12): 368–379.
Nishimura, G. (1954) Ultrasonic machining—Part I. Journal of Fracture Engineering, Tokyo University,
24(3): 65–100.
Pandey, P.C.; Shan, H.S. (1980) Modern Machining Processes, Tata McGraw-Hill, New Delhi, pp. 7–38.
Pei, Z.J., Ferreira, P.M., Haselkorn, M. (1995a) Plastic flow in ultrasonic machining of ceramics. J. Mate-
rials Processing Technology, 48(1–4): 771–777.
Pei, Z.J.; Khanna, N.; Ferreira, P.M. (1995b) Rotary ultrasonic machining of structural ceramics—A
review. Ceramic Engineering and Science Proceedings, 16(1): 259–278.
Pei, Z.J.; Prabhakar, D.; Ferreira, P.M.; Haselkorn, M. (1995c) A mechanistic approach to the prediction
of material removal rates in rotary ultrasonic machining. Journal of Engineering for Industry, 117:
142–151.
Pei, Z.J.; Ferreira, P.M. (1998) Modeling of ductile-mode material removal in rotary ultrasonic machin-
ing. International Journal of Machine tools and Manufacture, 38(10=11): 1399–1418.
Pentland, E.W.; Ektermanis, J.A. (1965) Improving ultrasonic machining rates—Some feasibility studies.
Journal of Engineering For Industry, Trans. of the ASME, 87: 39–46.
Perkins, J. (1972) An outline of power electronics. Ultrasonics, 1(5): 42–47.
Prabhakar, D.; Haselkorn, M. (1992) An experimental investigation of material removal rates in rotary
ultrasonic machining. Transactions of NAMRI=SME, 20: 211–218.
Rajurkar, K.P.; Wang, Z.Y. (1999) Micro removal of ceramic material in the precision ultrasonic machin-
ing. Precision Engineering, 23(2): 73–78.
Ramulu, M. (2005) Ultrasonic machining effects on the surface finish and strength of silicon carbide
ceramics. International Journal of Manufacturing Technology Management, 7(2=3=4): 107–125.
Rao, R.V.; Pawar, P.J.; Davim, J.P. (2010) Parameter optimization of ultrasonic machining process
using nontraditional optimization algorithms. Materials and Manufacturing Processes, 25(10):
1120–1130.
Richardson, D.W.; Robare, M.W. (1978) Turbine component machining development, Proc. Conference
on Ceramic Machining and Surface Finishing, Naval Research Lab, Gaithersburg, Md., 278–289.
Riddie, V. (1973) Cavitation erosion—A survey of the literature 1940–1970. Wear, 23: 133–137.
Robare, M.W.; Richerson, D.W. (1977) Proceedings of the ARPA=NAVSEA-Garrett=Ai Research Ceramic Gas
Turbine Engine Demonstration Program Review at Rotor Blade Machining Development, Marine Maritime
Academy, Castine, ME.
Ross, P.J. (1988) Taguchi Technique for Quality Engineering, McGraw-Hill Book Company, New York, pp. 45–48.
Rozenberg, L.D. (1973) Physical Principles of Ultrasonic Technology, Plenum Press, New York, (1–2): 20–53.
Rozenberg, L.D.; Kazantsev, V.F.; Makarov, L.O. (1964) Ultrasonic Cutting, Consultant Bureau, New York,
97–102.
Rutan, H.L. (1984) Ultrasonic machining (impact grinding), InProceedings of Topical Meetings on
Optical Fabrication and Test, U.S. Department of Energy.
Saha, J.; Bhattacharya, A.; Mishra, P.K. (1988) Estimation of material removal rates in USM process—A
theoretical and experimental study. Proc. 27th Int. Matador Conf., Manchester, England, 31–46.
Sahay, C.; Ghosh, S.; Kammila, H.K. (2011) Analysis of ultrasonic machining using Monte Carlo simula-
tion. Proceedings of the ASME 2011 International Mechanical Engineering Congress &
Exposition, November 11–17, 2011, Denver, Colorado, USA.
Satyanaryana, A.; Krishan Reddy, B.G. (1984) Design of velocity transformers for ultrasonic machining.
Electrical India, 24(14): 11–20.
Scab, K.H.W. (1990) Parametric studies of ultrasonic machining. SME Tech. Paper, MR90–294: 11.
Seah, K.H.W.; Wong, Y.S.; Lee, T.C. (1993) Design of tool holders for ultrasonic machining using FEM.
Journal of Materials Processing Technology, 37(1–4): 801–806.
Shaw, M.C. (1956) Ultrasonic grinding. Annals of CIRP, 5: 25–53.
Sharma, A.; Mishiro, S.; Suzuki, K.; Imai, T. (2003) A new longitudinal mode ultrasonic transducer with
an eccentric horn for micro machining. Key Engineering Materials, 238–239: 147–152.
378 J. Kumar

Sharman, R.C.; Bowen, P.; Aspinwall, D.K. (2001) Ultrasonic assisted turning of gamma titanium alumi-
nide, Proceedings of 13th International Symposium for Electromachining, Spain, Part I, 939–951.
Shen, J.; Pei, Z.J.; Lee, E.S. (2008) Support vector fuzzy adaptive network in the modeling of material
removal rate in rotary ultrasonic machining. Journal of Manufacturing Science and Engineering, 130:
1–8.
Singh, R.; Khamba, J.S. (2003a) Comparing the machining characteristics of titanium with cylindrical
and conical horn in ultrasonic drilling. Proceedings of the National Conference on Recent Developments
in Mechanical Engineering (NCME 2003), TIET Patiala, India.
———. (2003b) Investigating the machining characteristics of titanium alloy (Titan-31) using ultrasonic
drilling. Proceedings of the 13th National Conference of ISME 2003, IIT, Rorkee, India.
Singh, M.K. (2007) Unconventional Manufacturing Processes. New Age International Publishers, New
Delhi, pp. 4–8.
Singh, J.; Gill, S.S. (2009) Fuzzy modeling and simulation of ultrasonic drilling of Porcelain Ceramic with
Hollow Stainless Steel Tools. Materials and Manufacturing Processes, 24(4): 468–475.
———. (2006) Ultrasonic Machining of titanium and its alloys-a review. Journal of Materials Processing
Technology, 173: 127–131.
———. (2007a) Taguchi technique for modeling material removal rate in ultrasonic machining of
titanium. Materials Science and Engineering, 460–461: 365–369.
———. (2007b) Macro-model for ultrasonic machining of titanium and its alloys: designed experi-
ments. Journal of Engineering Manufacture, 221: 221–235.
———. (2008) Comparison of slurry effect on machining characteristics of titanium in ultrasonic drill-
ing. Journal of Materials Processing Technology, 197(1–3): 200–205.
Snoyes, R. (1986) Non-conventional machining techniques: The state of art. Advances in Non-Traditional
Machining, ASME, 1–20.
Spur, G.; Brucker, Th.; Holl, S.E. (1997) Ultrasonic machining of ceramics. Industrial Ceramics, 17(1):
29–34.
SONIC-MILL 500 W model (2002) Instruction Manual for Stationary Ultrasonic Machining, Albuquerque,
NM, USA.
Soundrajan, V.; Radhakrishnan, V. (1986) An experimental investigation on the basic mechanisms
involved in the ultrasonic machining. International Journal of Machine Tool Design and Research,
26(3): 307–321.
Sreejith, P.S.; Ngoi, B.K.A. (2001) Material removal mechanisms in precision machining of new materi-
als. International Journal of Machine Tools and Manufacture, 41: 1831–1843.
Tabatabaei, S.M.K.; Behbahani, S.; Mirian, S.M. (2013) Analysis of ultrasonic assisted machining
(UAM) on regenerative chatter in turning. Journal of Materials Processing Technology, 213:
418–425.
Thoe, T.B.; Aspinwall, D.K.; Wise, M.L.H. (1995) The effect of operating parameters on ultrasonic con-
tour machining. Proc. 12th Annual Conference of the Irish Manufacturing Committee, Cork, Ireland,
Sep., 1995, 305–312.
Thoe, T.B.; Aspinwall, D.K. (1999) Combined ultrasonic and electric discharge machining of ceramic
coated nickel alloy. Journal of Materials Processing Technology, 92–93: 323–328.
———. (1998) Review on ultrasonic machining. International Journal of Machine Tools Manufacture, 38(4):
239–255.
Treadwell, C.; Hu, P.; Zhang, J.M. (2002) Modeling of material removal rate in rotary ultrasonic machin-
ing: Designed experiments. Journal of Materials Processing Technology, 129(1–3): 339–344.
Tsutsumi, C.; Okano, K.; Suto, T. (1993) High quality machining of ceramics. Journal of of Materials Processing
Technology, 37: 639–650.
US 400 (1994) Ultrasonic machining system. Brochure from Erosonic AG.
Venkatesh, V.C. (1983) Machining of glass by impact processes. Journal of Mechanical Working Technology,
8: 247–260.
Wang, H.; Plebni, L.J.; Sathyanaryanan, G. (1998) Modeling considerations for ultrasonic machining.
Abrasives, Oct–Nov 1998, 8.
Wang, Q.; Cong, W.; Pei, Z.J.; Gao, H.; Kang, R. (2009) Rotary ultrasonic machining of dihydrogen phos-
phate (KDP) crystal: An experimental investigation on surface roughness. Journal of Manufacturing
Processes, 11: 66–73.
Ultrasonic Machining 379

Wang, Z.Y.; Rajurkar, K.P. (1995) Dynamic analysis of ultrasonic machining process. Proceedings of the
1995 ASME International Mechanical Engineering Congress and Exposition, Part I: 87–97.
Wansheng, Z.; Wang, Z.; Shichun, D.; Guanxin, C. (2002) Ultrasonic and electric discharge machining
to deep and small hole on titanium alloy. Journal of Materials Processing Technology, 120: 101–106.
Weller, E.J. (1984) Non-traditional machining processes. Society of Manufacturing Engineers, 15–71.
Wiercigroch, M.; Neilson, R.D.; Player, M.A. (1999) Material removal rate prediction for ultrasonic drill-
ing of hard materials using an impact oscillator approach. Physics Letters, 259: 91–96.
Willard, G.W. (1953) Ultrasonically induced cavitation. Journal of Acoustic Society of America, 25: 669.
Wu, J.; Cong, W.; Williams, R.E.; Pei, Z.J. (2011) Dynamic process modelling for rotary ultrasonic
machining of alumina. Journal of Manufacturing Science and Engineering, 133.
Xu, W.L.; Han, L. (1999) Piezoelectric actuator based active error compensation of precision machin-
ing. Measurement Science and Technology, 10(2): 106–111.
Ya, G.; Quin, H.W.; Yang, S.C. (2002) Analysis of rotary ultrasonic machining mechanism. Journal of
Materials Processing Technology, 129(1–3): 182–185.
Yan, C.; Biing, H. (2001) Surface modification of Al-Zn-Mg alloy using the combined process of EDM
and USM. Journal of Materials Processing Technology, 115(3): 359–366.
Zarepour, H.; Yeo, S.H. (2012) Single abrasive particle impingements as a benchmark to determine
material removal modes in micro ultrasonic machining. Wear, 288: 1–8.
Zeng, W.M.; Li, Z.C.; Pei, Z.J.; Treadwell, C. (2005) Experimental observation of tool wear in rotary ultra-
sonic machining of advanced ceramics. International Journal of Machine Tools and Manufacture, 45:
1468–1473.
———. (1997) Study of a new kind of combined machining technology of ultrasonic machining and elec-
trical discharge machining. International Journal of Machine Tools and Manufacture, 37(2): 193–199.
———. (1999) Material removal rate Analysis in the ultrasonic machining of engineering ceramics.
Journal of Materials Processing Technology, 88(1): 180–184.
Zhinxin, J.; Zhang, J.H.; Xing, A. (1995) Combined machining of USM and EDM for advanced
ceramics. Journal of Advanced Materials, 26(3): 1620.

You might also like