You are on page 1of 17

Int. J. Mach. Tools Manufact. Vol. 38, No. 4, pp.

239–255, 1998
 1998 Elsevier Science Ltd. All rights reserved
Pergamon Printed in Great Britain
0890–6955/98$19.00 + .00

PII: S0890–6955(97)00036-9

REVIEW ON ULTRASONIC MACHINING


T. B. THOE,† D. K. ASPINWALL†‡§ and M. L. H. WISE‡
(Received 23 January 1996; in final form 30 April 1997)

Abstract—Ultrasonic machining is of particular interest for the cutting of non-conductive, brittle workpiece
materials such as engineering ceramics. Unlike other non-traditional processes such as laser beam, and electrical
discharge machining, etc., ultrasonic machining does not thermally damage the workpiece or appear to introduce
significant levels of residual stress, which is important for the survival of brittle materials in service. The
fundamental principles of ultrasonic machining, the material removal mechanisms involved and the effect of
operating parameters on material removal rate, tool wear rate and workpiece accuracy are reviewed, with
particular emphasis on the machining of engineering ceramics. The problems of producing complex 3-D shapes
in ceramics are outlined.  1998 Elsevier Science Ltd. All rights reserved

1. AN OVERVIEW OF ULTRASONIC MACHINING AND APPLICATIONS

Ultrasonic machining (USM) is a non-conventional mechanical material removal process


generally associated with low material removal rates, however its application is not limited
by the electrical or chemical characteristics of the workpiece materials. It is used for
machining both conductive and non-metallic materials; preferably those with low ductility
[1–5] and a hardness above 40 HRC [6–12], e.g. inorganic glasses, silicon nitride,
nickel/titanium alloys, etc. [13–24]. Holes as small as 76 ␮m in diameter can be machined
[25], however, the depth to diameter ratio is limited to about 3:1 [8, 12].
The history of ultrasonic machining (USM) began with a paper by R. W. Wood and
A. L. Loomis in 1927 [26, 27] and the first patent was granted to L. Balamuth in 1945 [7,
28, 29]. USM has been variously termed ultrasonic drilling; ultrasonic cutting, ultrasonic
dimensional machining; ultrasonic abrasive machining and slurry drilling. however, from
the early 1950s it was commonly known either as ultrasonic impact grinding or USM [8,
25, 28, 30, 31].
In USM, high frequency electrical energy is converted into mechanical vibrations via
a transducer/booster combination which are then transmitted through an energy focusing
device, i.e. horn/tool assembly [1, 17, 18, 30, 32]. This causes the tool to vibrate along
its longitudinal axis at high frequency (usually ⱖ 20 kHz) with an amplitude of 5–50 ␮m
[16, 33, 34]. Typical power ratings range from 50–3000 W [35] and can reach 4 kW in
some machines. A controlled static load is applied to the tool and an abrasive slurry
(comprising a mixture of abrasive material; e.g. silicon carbide, boron carbide, etc. sus-
pended in water or oil) is pumped around the cutting zone. The vibration of the tool causes
the abrasive particles held in the slurry between the tool and the workpiece, to impact the
workpiece surface causing material removal by microchipping [19]. Fig. 1 shows the basic
elements of an USM set up using either a magnetostrictive or piezoelectric transducer
with brazed and screwed tooling.
Variations on this basic configuration include:–
쐌 Rotary ultrasonic machining (RUM). Here the tool is excited and simultaneously rotated
so reducing out-of-roundness to about 1/3 the values obtained in conventional USM

†School of Manufacturing and Mechanical Engineering, University of Birmingham, Edgbaston, Birmingham,


B15 2TT, U. K.
‡Interdisciplinary Research Centre in Materials for High Performance Applications, University of Birmingham,
B15 2TT, U. K.
§Author to whom correspondence should be addressed

239
240 T. B. Thoe et al.

Static Load
Horn Oscillator
Input Stroke (High Frequency Power Support
Transducer
Input Stroke
Loop
Water Cooling Jacket

Laminated Nickel
λ/2 Generator Node Magnetostrictive Transducer

Piezoelectric Transducer
Loop Silver Soldered/Brazed Joint

Booster Horn
λ/2 First Node Resonant Support
Mechanical
Amplifier Horn
Stub
Loop Tool Vibration
Silver Soldered/Brazed Joint
Tool

λ/2 Second Node


Mechanical Abrasive Slurry
Amplifier Workpiece
Screw Threaded Tool
Loop

Fixture
Verticle Stroke Output
Of Tool Slurry tank Slurry Pump

Fig. 1. Basic elements of USM head [28, 36–40].

[41]. Typical rotational speeds are ⬇ 300 rpm, but with diamond impregnated tools,
rotational speeds can be as high as 5000 rpm.
쐌 USM combined with electrical discharge machining (EDM) [16, 18, 20, 26, 34, 36].
쐌 Ultrasonic assisted conventional/non-conventional machining. USM assisted turning is
claimed to reduce machining time, workpiece residual stresses and strain hardening,
and improve workpiece surface quality and tool life compared to conventional turning
[12, 36, 42–45].
쐌 There are also non-machining ultrasonic applications such as cleaning, plastic/metal
welding, chemical processing, coating and metal forming.

1.1. Contour machining using ultrasonic techniques


Many USM applications are involved in drilling where a tool of either simple or com-
plex cross section penetrates axially into the workpiece, to produce either a through or
blind hole of the required dimensions. Where a three dimensional cavity is required (i.e.
one in which the depth varies), a process analogous to die sinking is generally employed
[7, 10, 27, 37, 43, 46–48], see Fig. 2. Using this technique graphite electrodes for EDM
have been shaped in 30 minutes instead of the 20 hours required by copy milling [2, 4,
49, 50]. The problem with using tools of complex form, however, is that they are not
subject to the same machining rate over the whole of their working surface and experience
differential wear rates, both of which affect the product shape [51]. In addition, there are
also greater problems in tuning a complex tool to achieve maximum performance com-
pared to more basic tools.
An alternative approach is to use a simple “pencil” tool and contour machine the com-
plex shape with a CNC programme, see Fig. 3. Recently, the feasibility of using this
technique has become of interest and has been investigated in a number of countries
including the UK, France, Switzerland, Japan, etc. [26, 52]. A few CNC controlled path
rotary USM systems are available commercially such as the SoneX 300 from Extrude
Hone Limited (France); and the Erosonic US400/US800 from Erosonic AG (Switzerland).
Review on ultrasonic machining 241

Fig. 2. Silicon nitride turbine blade counter-sunk using USM [7].

Fig. 3. Carbon fibre composite acceleration lever, holes and outline profile machined by USM [7].

1.2. Ultrasonic machining of ceramic materials


Advanced ceramics are increasingly being used for applications in the aerospace, auto-
motive and electronics sectors of industry. They offer a number of advantages over metals
for automotive valves and cylinder sleeves [14, 15] and their good chemical stability and
high temperature performance offers the possibility of greater thermodynamic efficiency
in gas turbine applications [30, 53, 54]. Sintered alumina, silicon carbide and silicon nitride
products generally have a hardness ⬎ 1500 Hv and therefore diamond grinding is usually
the only feasible/economic method of machining components to final shape. While this
may be acceptable and perfectly feasible with components that comprise cylindrical, flat
and curved surfaces, problems can arise where the surface to be machined has a more
complex topography or where in-service operating characteristics dictate a particular work-
piece surface integrity.
The majority of engineering ceramics are electrical insulators and although this can be
an advantage in terms of function, it is a significant disadvantage in relation to the machin-
ing of components such as ceramic or ceramic coated turbine blades. Equivalent metal
products rely a great deal on the use of non-conventional processes, such as electro-chemi-
cal machining (ECM) and EDM. The former is used extensively for the production of
aerofoil sections while the latter is used for the machining of blade cooling holes. Unfortu-
nately, both processes rely on the fact that the workpiece material is conductive. In the
case of EDM, the workpiece needs to have an electrical resistivity of less than 100 ⍀cm
[16, 26, 33, 55].
With any safety critical component such as a turbine blade, which is highly stressed
during operation, workpiece surface integrity is a key feature and particularly so where
the component is made of ceramic or employs a ceramic coating. The machining process
242 T. B. Thoe et al.

used should impart as little damage as possible to the finished surface/sub-surface. All
conventional cutting operations such as turning or grinding will to a lesser or greater
degree cause some type of surface damage [24, 48, 56]. This is also true of non-conven-
tional processes, such as EDM or laser beam machining (LBM) which rely on a thermal
cutting mechanism. For example, EDM can cause a thermally altered surface zone up to
50 ␮m deep with microcracking [55]. In contrast, ultrasonic machining (USM) is a non-
thermal process which does not rely on a conductive workpiece and is suited to the machin-
ing of ceramic materials. The process produces little or no surface/sub-surface damage or
impose particular stress regimes. Quoted dimensional accuracy is ± 5 ␮m [36] and surface
finishes of Ra 0.51–0.76 ␮m can be achieved [37].
USM data on graphite, silicon carbide and a range of ceramic materials from work by
Gilmore [6], Kremer [33] is summarised in Table 1.

2. BASIC ELEMENTS OF AN ULTRASONIC MACHINE TOOL

2.1. The ultrasonic generator and ultrasonic transducer


With a conventional generator system, the horn and tool are set up and mechanically
tuned by adjusting their dimensions to achieve resonance. Recently, however, resonance
following generators have become available which automatically adjust the output high
frequency to match the exact resonant frequency of the horn/tool assembly [6]. They can
also accommodate any small errors in set up and tool wear, giving minimum acoustic
energy loss and very small heat generation [33]. The power supplied depends on the size
of the transducer [35]. Some generators are designed with safety features such as an auto-
matic switch off in cases of horn fracture, horn/tool joint failure, etc. [17, 31, 33].
Transducers vibrate in a longitudinal or compressive mode. For industrial applications,
either a magnetostrictive [12, 26, 40, 57] or a piezoelectric device [35, 39] is used. Because
of its lower Q value (Q is a measurement of the sharpness of the peak value of the energy),
the magnetostrictive transducer allows vibration to be transmitted over a wide frequency
band (for example, 17–23 kHz for a 20 kHz generator) [58]. It also allows greater horn
design flexibility and can accommodate tool wear. In addition, the horn can also be
redesigned/machined several times without critical loss of amplitude [4, 30, 46]. The major
drawback of magnetostrictive transducers are their high electrical losses (e.g. eddy current
loss) and low energy efficiencies ( ⬇ 55%) [40]. These losses appear as heat, therefore,
the transducer has to be air/water cooled and the size of the transducer is bulky. Also,
the transducer is not capable of generating high vibration intensities as compared to piezoe-
lectric types [59, 60]. A typical piezoelectric transducer [26, 42, 53, 61] consists of two
discs of lead zirconate-titanate, or other synthetic ceramic [62] with a thickness usually
less than 10% of the total ultrasonic transducer length [63]. Piezoelectric transducers have
high energy efficiencies ( ⬇ 90–96%) and consequently do not require any cooling [18,
28, 59]. They are not liable to heat damage and appear to be more easily adaptable for
rotary operations [61] and more easily constructed.

Table 1. Data from various materials ultrasonically machined using 320 mesh abrasive[6, 33].

Workpiece Hardness Hv Surface Recommended MRR(mm3/min) MRR(mm3/min)


Materials Roughness Ra Abrasive 5 mm ⭋ Tool 10 mm ⭋ Tool
(␮m)

Graphite 65 1–2 SiC/B4C 164 224


Silicon Oxide 500 0.85 SiC/B4C 39 50
Aluminium 1000 0.9 SiC/B4C 7.6 9.3
Oxide
Zirconia 1100 0.75 B4C 0.65 3.1
Sialon 1500 0.4 B4C 1.2 1.8
Silicon Carbide 2400 0.3 B4C 0.6 3.5
Review on ultrasonic machining 243

2.2. The ultrasonic horn and tool assembly


The horn is variously referred to as an acoustic coupler, velocity/mechanical trans-
former, tool holder, concentrator, stub or sonotrode, see Fig. 4. The oscillation amplitude
at the face of the transducer is too small (0.001–0.1 ␮m) [23, 40, 64] to achieve any
reasonable cutting rate, therefore, the horn is used as an amplification device [26, 60].
Optimum tuning is specific for each horn material [65] which should possess a high mech-
anical Q, good soldering and brazing characteristics, good acoustic transmission properties
and high fatigue resistance at high working amplitude. It should also be corrosion resistant
and strong enough to take screw attachments. Monel, titanium 6-4 (IMI 318), AISI 304
stainless steel, aluminium and aluminium bronze are commonly used [1, 4, 20, 40, 64–67].
The tool should be so designed to provide the maximum amplitude of vibration at the
free end (antinode) at a given frequency [61]. The material used should have high wear
resistance, good elastic and fatigue strength properties, and have optimum values of tough-
ness and hardness for the application [16, 27, 64]. Tungsten carbide, silver steel, and
Monel are commonly used tool materials. Polycrystalline diamond (PCD) has recently
been detailed for the machining of very hard workpiece material such as hot isostatically
pressed silicon nitride [68].
Tools can be attached to the horn by either soldering or brazing, screw/taper fitting.
Alternatively, the actual tool configuration can be machined on to the end of the horn [1,
4, 27, 35, 66, 70, 71]. Threaded joints are conventionally used because of quick and easy
tool changing, however, problems can occur such as self-loosening, loss of acoustic power,
fatigue failure, etc [72]. When deep hole drilling with a trepanning tool, the ability to feed
the abrasive through the centre of the horn and tool is a great advantage thus reducing
side friction [27].

2.3. The tool feed advancing mechanism and the abrasive feed system
The tool is normally held against the workpiece by a static load exerted via a
counterweight/static weight, spring, pneumatic/hydraulic or solenoid feed system [16, 26,
27, 40, 73]. For optimum results, the system should maintain a uniform working force
while machining and be sufficiently sensitive to overcome the resistance due to the cutting
action [16, 40]. The applied force must be chosen carefully since too low a setting will
not yield the maximum cutting rate whereas too high a setting will cause jamming between
the tool and abrasive [3, 70]. Static load values of about 0.1–30 N are typically used. The
force is particularly critical when drilling small holes less than 0.5 mm diameter as bending
of the tool can occur under too high a load.
The slurry is usually pumped across the tool face by jet flow, suction, or a combination
of both as shown in Fig. 5 [13, 16, 28, 40, 74]. It acts as a coolant for the horn, tool and
workpiece, supplies fresh abrasive to the cutting zone and removes debris from the cutting
area [2, 25, 27, 28]. The slurry also provides a good acoustic bond between the tool,

Fig. 4. Various horn designs with and without additional tool heads [69].
244 T. B. Thoe et al.

Fig. 5. Methods of slurry delivery [13, 16, 28, 40, 74].

abrasive and workpiece, allowing efficient energy transfer. The infeed tube is connected
at, or adjacent to, a nodal plane of the horn to avoid damping effects [16, 27]. The most
common abrasive materials used are aluminium oxide, silicon carbide, boron carbide, etc.
[4, 12, 24, 27, 37, 75–78]. 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, water meets
most of these requirements [3, 26, 28].
3. MATERIAL REMOVAL MECHANISMS

Extensive work on the mechanism of material removal has been done by Shaw [35],
Miller [79], Cook [80], Rozenberg et al. [7] and others [22, 23, 43, 60]. These mechanisms
are detailed in Fig. 6 and comprise:-
쐌 Mechanical abrasion by direct hammering of the abrasive particles against the workpiece
surface [10, 28, 34, 35, 37, 40, 50, 60, 70, 81];
쐌 Micro chipping by impact of the free moving abrasive particles [28, 35, 37, 50, 70,
81, 82];
쐌 Cavitation effects from the abrasive slurry [10, 27, 35, 48, 50, 82];
쐌 Chemical action associated with the fluid employed [28, 35].
The individual or combined effect of the above mechanisms result in workpiece material
removal by shear [13, 36, 70], by fracture (for hard or work hardened material) [13] and
displacement of material at the surface, without removal, by plastic deformation [13, 36]
which will occur simultaneously at the transient surface.
With porous materials like graphite as opposed to hardened steels and ceramics, cavi-
tation erosion is a significant contributor to material removal [10, 23, 28, 35, 37, 81].
Markov [21] and others [27, 35] considered that cavitation erosion and chemical effects
were of secondary significance with the majority of workpiece materials acting essentially
to weaken the workpiece surface, assist the circulation of the abrasive and the removal
of debris. In RUM, Komaraiah et al. [83] and Enomoto [84] found that the theoretical
static load required for the formation of Hertzian cracks in brittle materials was less than
that for sliding indentation.
Static Load & Vibration Slurry
[H2O+Abrasives (e.g. B4C)]

Tool bottom Bottom Wear

Workpiece
Frontal Wear

Localised Hammering Free Impact Cavitation Erosion

Fig. 6. USM material removal mechanisms [81].


Review on ultrasonic machining 245

3.1. Effect of various operating parameters on material removal rate (MRR)


The amplitude, ␰, of the ultrasonic tool before machining can be measured by using
either an accelerometer [10], an eddy current probe [30, 85], a laser dopplermeter [86] or
laser speckle pattern interferometer [58]. A higher amplitude is obtained by using a tool
with a high transformation ratio, i.e. the ratio of transducer/tool diameter [27, 39]. Ideally,
the amplitude should be equal to the mean diameter of the abrasive grit used in order to
optimise the cutting rate [3, 4, 6, 10, 13, 24, 49, 66]. Shaw [35] showed that MRR ⬀ ␰3/4
while other researchers [77, 79, 87] have advocated that MRR ⬀ ␰, and yet others [7, 13,
27, 28, 74, 75, 88, 89] have suggested that MRR ⬀ ␰2 at constant frequency and static
load. In general, MRR increases as the amplitude of tool vibration increases (everything
else remaining constant) [40, 77], even so, there exists an amplitude level at which MRR
decreases, as shown in Fig. 7.
Some authors [6, 27, 75] have suggested that at constant amplitude, MRR ⬀ f 2 up to
a frequency, f, of 400 Hz. At higher frequencies, up to 5 kHz, a linear relationship between
frequency and MRR is found. Above an upper threshold value, the MRR falls off rapidly
where MRR⬀√f [3, 75]. Rozenberg et al. [7] and Kainth et al. [22] have shown that, in
practice, an increase in static load from zero, with other parameters constant, yields an
approximately linear relationship between MRR and static load. Above an optimum value,
the MRR decreases owing to a reduction in the size of the abrasive grains reaching the
tool/workpiece interface and insufficient slurry circulation [7, 16, 26, 43, 60, 75, 88, 90].
The optimum static load for the maximum machining rate has been found to be dependent
on the tool configuration (e.g. cross-sectional area and shape), as shown in Fig. 8, the
amplitude and mean grit size [16, 23, 27, 70, 91]. Kops [92] indicated that the use of a
smaller than optimal value (based on MRR) for the static load, is better for reducing
abrasive wear and increasing tool life.
The abrasive material should be harder than the workpiece and, typically, larger abrasive
grit sizes [2, 5, 19, 27, 35, 70, 93] and higher slurry concentrations [3, 13, 23, 28, 59,
66, 81] yield higher MRR. On increasing the abrasive grit size or slurry concentration,
an optimum MRR is reached. Any further increase in either aspect results in difficulty in
the larger grains reaching the cutting zone [5, 7, 10, 40, 60, 87] and a consequent fall in
MRR. A slurry concentration of 30% has been recommended [1, 13, 27, 59, 77, 87, 94].
Kazantsev [74] claimed that 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
the suction pumping system, it yielded a 2–3 times higher MRR [23]. In terms of MRR,

1 2 3 4

15 13 9 3
17 17 17 17
Tool cross-sections

ν [mm/min] 1

30

20 3

10
0 10 20 30 40 ξ [µm]

Fig. 7. Relationship between penetration rate and oscillation amplitude [74].


246 T. B. Thoe et al.

0.020
[0.51mm]

0.015

Penetration per minute, in


Tool section
1/4 in
[6.35mm]

0.010
Tool section 1-1/2 in
[38mm]

1/2 in Tool
0.005 [12.7mm] Section

0
0 2 4 6 8 10
[3.64 Kg]
Static load, lb

Fig. 8. Penetration rate vs. static load for various areas [3, 23, 37, 43, 70, 74, 88].

water is generally superior to most oils, benzene and glycerol-water mixtures [3, 26, 28].
Pentland et al. [13] and others [40, 43, 50, 95] found that by improving slurry circulation,
cavitation, contamination and blockage effects can be reduced or overcome.
Although USM can be used on both hard and soft workpiece materials, brittle materials
are preferable [12, 27, 70]. Harder materials are cut by brittle fracture whereas softer,
ductile materials like mild steel are cut by plastic shear. In such cases, the abrasive grit
tends to embed itself in the workpiece [12, 13, 28, 36]. Low workpiece fracture toughness
or increasing the ratio of tool hardness to Young’s modulus (H/E) will result in a higher
MMR [25, 41, 55, 84] as shown in Figs 9 and 10. The mechanical properties of a work-
piece material and its fracture behaviour are important in terms of machinability [3, 41, 43].
It has been reported that the cutting rate is directly proportional to the tool form [95]
and shape factor (ratio of tool perimeter to tool area) [16, 60, 77, 79]. The tool form defines
the resistance to slurry circulation: a tool of narrow rectangular cross section yielding a
higher machining rate than one with a square cross-section of the same area [16, 40, 87]
as shown in Fig. 11. Goetze [77] reported that for tools with equal contact areas, there is
an increase in penetration rate for tools with larger perimeters; the effect being mainly
due to the difficulty of adequately distributing the abrasive slurry over the machining zone
[3, 16, 40, 96].

350 Glass 35
Material Removal Rate
300 30
Relative Wear
Y.PSZ
250 25
Relative Wear [%]
MRR [mm3/min]

200 HIPSN 20
AL2O3

150 HPSN 15
SSiC
100 SiSiC [Refel] 10
RBSN

50 5

HPSiC SSiC
0 0
0 2 4 6 8 10 12

Fracture Toughness [MPam1/2]

Fig. 9. Effect of fracture toughness (KIC) for a range of ceramics on USM MRR and relative wear [55].
Review on ultrasonic machining 247

Conventional Porcelain
50 USM
45
40
Rotary USM
35

MRR (mm3/min)
30
25
20
15 Ferrite Glass
10 Alumina
5 Carbide

0
1.28 2.36 3.87 8.1 13.26 (×10–6)
H/E

Fig. 10. Effect of H/E on MRR for various materials under USM & RUM [25, 41].

0.125

1.265 mm
0.100
Rate of machining, mm min–1

12.65 mm

0.075

4 mm
0.050

4 mm
0.025

0
0 4 8 12 16
Pressure, Nm–2

Fig. 11. Effect of tool shape on MRR for tools with same cross-sectional area [16, 40].

For tools with a small cross-sectional area, the adjustment of the static load for optimum
working becomes more critical [3], but cutting rates tend to be greater under otherwise
identical conditions (see Fig. 8). Several researchers [70, 71, 77, 90] have concluded that
the optimum internal to external diameter ratio for a trepanning tool is approximately 0.45.
A lower limit on the thickness of the tool has been suggested of not less than five times
the abrasive grit size [16, 27]. The hardness of the tool material influences the MRR, the
tool wear rate and component accuracy [41]. Komaraiah et al. [97] and others [40, 86]
have shown that tool materials can be ranked in order of superiority as follows: Nimonic
80A ⬎ thoriated tungsten ⬎ silver steel ⬎ stainless steel ⬎ maraging steel ⬎ titanium
⬎ mild steel. Neppiras [27] using other tool materials gave the following ranking: tungsten
carbide ⬎ brass ⬎ mild steel ⬎ silver steel ⬎ stainless steel ⬎ copper. Tools with diamond
tips have been shown to have good material removal characteristics and extremely low
wear rates [6].
In RUM, the rotary motion of the tool enhances MRR, workpiece accuracy and in some
cases reduces cutting forces [12, 76] and increases tool life [27]. MRR in RUM is nearly
6–10 times higher than when diamond surface grinding under similar experimental con-
ditions [66] and up to 4 times that in standard USM [90]. Komaraiah et al. [83] stated
that the superior performance of RUM over standard USM may be explained by the com-
bined effects of indentation of the workpiece surface, sliding contact between the embed-
248 T. B. Thoe et al.

ded grains on the tool/workpiece and rolling contact between the free abrasive
grains/workpiece. In RUM, Prabliakar et al. [66] and Komaraiah et al. [90] showed that
higher rotational speeds resulted in higher MRR. In general, the optimum drilling depth
should be 2–5 times the diameter of the tool, but this is governed by the slurry supply
conditions [16, 87].
4. TOOL WEAR

Tool wear is an important variable in USM, affecting both MRR and hole accuracy
[38, 28, 87, 94, 98]. The complex tool wear pattern in USM can be divided into longitudi-
nal wear, WL [71, 87, 94], and lateral/side/diametral wear, WD [99], some of which will
occur as a result of cavitation or suction wear [38, 71, 75, 100].

4.1. Effect of various operating parameters on tool wear


Adithan [71] and Venkatesh [38] reported that tool wear is a maximum at a particular
static load which may be considered optimum for the point of view of maximum MRR.
The MRR decreases beyond this optimum static load. Tool wear tends to increase when
harder and coarser abrasives are used, see Fig. 12. As a consequence, harder abrasives,
like boron carbide, cause higher tool wear than softer abrasives like silicon carbide for a
tool of the same cross-sectional area [38, 99]. Tool wear is affected by workpiece hardness
and can also be affected by the toughness of the workpiece. Transformation toughened
ceramics show poorer USM behaviour, giving relatively high tool wear [55]. They also
tend to wear the abrasive at a higher rate than when machining more conventional ceramics
[6, 101].
If the hardness of the tool increases by work hardening, the penetration of the abrasive
grains into the tool will decrease resulting in higher workpiece MRR. In addition, material
removal from the periphery of the work zone will be greater so that a convex surface will
be formed in the workpiece. This causes plastic deformation of the centre of the tool face,
forming a dish. It has also been found that the degree of hardening is highest at the
periphery and lowest at the centre for all tool materials [97]. As a result, soft materials,
e.g. copper and brass, are unsuitable as tools since they develop burrs at large oscillatory
amplitudes [3, 30]. They are also acoustically poor and attenuate the stress wave in large

D=6mm d=4mm 600grit

D=8mm d=6mm 600grit

D=6mm d=6mm 280grit

D=8mm d=6mm 280grit

0.7 Workpiece Material = Glass


Tool Material = Mild Steel
0.6 Abrasive = Boron Carbide

0.5
Total Wear [mm]

0.4

0.3

0.2

0.1

0
0 5 10 15 20

Total Time Of Drilling [min]

Fig. 12. Influence of abrasive grit size on tool wear [38, 99].
Review on ultrasonic machining 249

tools. The use of hard metals such as tungsten carbide reduces plastic deformation and
the wear of the tool surfaces [48].
To decrease WL, a material with a high value of the product of hardness, H, and impact
strength, Ki, (e.g. Nimonic 80A) is recommended as shown in Fig. 13. H and not Ki
affects WD significantly [97]. In an overall assessment of tool materials which exhibit
good resistance against WL and WD, Nimonic 80A, and thoriated tungsten or silver steel
have been recommended [3, 97]. Tool wear increases linearly with increasing depth of
hole drilled [71] and cutting time [10, 11].
5. THE EFFECT OF USM ON WORKPIECE SURFACE FINISH/ACCURACY

USM does not generate significant heating which might otherwise lead to the develop-
ment of a thermally damaged layer/zone or residual stress. Abrasive grain size has a sig-
nificant influence on workpiece accuracy and surface finish [4, 23, 26, 36, 40, 73, 82, 94].
A decrease in abrasive grain size during USM leads to lower Ra values, see Fig. 14. In
addition, the accuracy of the machined hole is improved [3, 10, 13, 28, 35, 41, 59, 66,
70, 75, 81] and a better surface finish is obtained on the bottom face than on the walls
of the cavity [2, 5, 7, 60, 86, 102]. Dam et al. [103] suggested that better surface finish
is obtained when feed rates and depths of cut are decreased.

10
Mild steel Titanium Maraging

Stainless steel Silver steel Nimonic 80A


Longitudinal Wear [mm]

1 Conventional USM

Rotary USM

0.1
1000 10000 100000

Hardness x Impact Strength [H.Ki]

Fig. 13. Effect of the product of hardness and impact strength on longitudinal tool wear [97].

3.0

2.5
Surface roughness, µm

2.0

1.5

1.0

(40 µin)
0.5

0
100 180 280 360 500 700 900
140 240 320 420 600 800 1000
Grit size

Fig. 14. Effect of surface roughness vs. grit size for boron carbide. Workpiece materials key: x = glass, 䊊 =
silicon semi-conductor; 왕 = mineralo-ceramic; 䊐 = hard alloy steel [16, 25, 37, 40, 41, 82].
250 T. B. Thoe et al.

Methods of improving workpiece wall surface finish have been suggested [16, 23, 27,
86, 87]. Kennedy et al. [16] and Koval’chenko et al. [5] pointed out the difficulty of
machining a flat at the bottom of a hole flat because of uneven slurry distribution across
the machining face, resulting in fewer active grits at the tool’s centre, Where the workpiece
is a hard ceramic, a slightly better surface finishes can be obtained than with a material
of lower hardness, values as low as 0.4 ␮m Ra are obtainable [2].
Production accuracy with USM holes must take account of both dimensional accuracy
(oversize) and form accuracy (out-of-roundness and conicity) [73]. The oversize is greatest
at entry, increasing rapidly with the depth-of-cut to a value which corresponds approxi-
mately to the upper limit of the abrasive grit size. An increase in the diameter-length ratio
increases lateral vibrations causing greater oversize [40, 50]. Shaw [35] and others [3, 23,
40, 73, 83] have shown that surface roughness improves with increased static load which
reduces the abrasive size and suppresses lateral vibrations of the tool, so minimising the
oversize/conicity/out-of-roundness of the holes produced as detailed in Fig. 15 [41].
Adithan et al. [73] found that the oversize with rectangular holes was greater than that
obtained with circular tools. Fig. 16 shows that workpiece materials with a high ratio of
hardness to Young’s modulus (H/E) are more susceptible to greater out-of-roundness.
Other factors affecting the dimensional or form accuracy are those relating to the pre-
cision of the acoustic elements and the USM tool [16, 27, 40, 73]. Conicity can be reduced
by using tungsten carbide and stainless steel as tool materials [73], an internal slurry
delivery system [23, 73, 86], tools with negative tapering walls or fine abrasives [16, 23,
27, 40, 87]. An increase in tool amplitude [40, 87] or the use of coarser abrasive grains
[81] while increasing the depth of penetration into the surface of workpiece, produces

0.2
Conventional USM
Out-of-roundness [mm]

0.15

0.1

Rotary USM (200 rpm)


0.05

0
0 0.25 0.5 0.75 1 1.25 1.50 1.75
Static Load [kgf]

Fig. 15. Effect of static load on out-of-roundness [41].

0.125 Conventional Porcelain


USM
Out-of-roundness (mm)

0.1
Rotary USM
Glass
0.075
Ferrite
Alumina
0.05

Carbide
0.025

0
1.28 2.36 3.87 8.10 13.26 (×10–6)
H/E

Fig. 16. Effect of H/E on out-of-roundness and MRR. Slurry: SiC of mesh size 180 and rotational speed of 200
rpm [41].
Review on ultrasonic machining 251

higher surface roughness. Kremer et al. [23] found that the USM of graphite resulted in
poor surface finish due to cavitation, contamination and graphite debris blockage effects.
Markov [87] concluded that Surface Roughness ⬀ MRR while Komaraiah et al. [41]
showed that Surface Roughness ⬀ (H/E)n for various materials under both standard USM
and RUM as shown in Fig. 17. The use of machine oil in place of water in a finishing
operation was found to improve the workpiece’s surface finish, but caused a reduction in
the cutting rate [16, 23, 27, 40, 87]. Where high workpiece accuracy is required, machining
must always be carried out in several stages. Dimensional accuracy of the order of ± 5 ␮m
can be obtained in most materials [36]. As shown in Fig. 14, the finer the abrasive grit
size, the finer the workpiece surface roughness [40].
6. HORN AND TOOL DESIGN

The theory and art of designing horns has been reviewed by several authors, but it is
not as yet fully understood [7, 63, 67, 104–106]. Traditional methods of acoustic horn
design are based on a differential equation which considers the equilibrium of an infini-
tesimal element under the action of elastic and inertia forces, which is then integrated
over the horn length to achieve resonance [56, 106, 107]. The horn length depends on the
working frequency and has no effect on energy amplification. Typical designs include
cylindrical, stepped, conical and exponential types [11, 88, 107]. Tuning is normally done
from the transducer (screw thread end) end where a tuning allowance of 10–15 mm should
be made [16, 25, 40, 65, 107, 108]. However, in the case of a booster horn, any reduction
in length must be done so that both ends are reduced equally to maintain its correct
resonance. Non-compliance with this rule will cause the nodal point to shift from the
point where the transducer was formerly supported and this will induce unwanted stresses,
eventually leading to the failure of the horn.
Recently, finite element modelling (FEM) has been used [11, 20, 33, 106, 107] to design
axisymmetric horn shapes. The analysis can take into consideration the weight of the tool,
any internal holes for slurry pumping and the method of fixing it to the transducer at the
required resonant frequency. FEM has also been used to assess the working stresses to
ensure safe stress limits [67]. Dam et al. [103] claimed that a horn can be designed which
converts the longitudinal ultrasonic action into a mixed lateral and longitudinal vibration
mode as shown in Fig. 18. This lateral motion obviously aids contouring work.
The uniformity of the horn vibration limits the process to the cutting of small shapes
typically less than 100 mm diameter [53]. An excess tool length of 2–3 mm generally
lowers the resonant frequency by about 0.5–1 kHz [20, 107], but when drilling very deep
holes this loss of efficiency may be overcome by making the tool itself resonate [27]. A
tool length to diameter ratio of less than 20:1 is generally recommended [12], but if the
tool length is greater than 10 mm, the horn must be shortened by an amount equivalent

Conventional Porcelain
2 USM
1.8
Glass
1.6
Surface Roughness (µm)

Rotary USM
1.4
1.2
1 Ferrite
0.8 Alumina

0.6 Carbide

0.4
0.2
0
1.28 2.36 3.87 8.1 13.26 (×10–6)
H/E

Fig. 17. Effect of H/E on surface roughness for various materials under USM & RUM [25, 41].
252 T. B. Thoe et al.

Ultrasonic
Input

Ultrasonic
Output

Diamond Coating

Fig. 18. Mix lateral and longitudinal vibration mode horn designed according to tuning fork principle [103].

to the weight of the tool [11]. Detailed guidelines for tool design are described by Rozen-
berg et al. [7].
7. CONCLUSIONS

1. USM is a non-thermal process which does not rely on a conductive workpiece and is
preferable for machining workpieces with low ductility and hardness above 40 HRC.
2. USM is believed to be a stress and damage free process.
3. For contour USM a resonance following generator is recommended because it can
automatically adjust the output high frequency to match the exact resonant frequency
of the horn/tool assembly. Such a generator can also accommodate any small errors in
set up and tool wear, giving minimum acoustic energy loss and very small heat gener-
ation.
4. The production of complex shapes may be achieved by sculpting using a simple tool
shape and CNC rather than die sinking using complex form tools, however, work in
this area is at an early stage.
5. Horn materials should possess a high mechanical Q, good soldering and brazing charac-
teristics, good acoustic transmission properties and high fatigue resistance at high work-
ing amplitude. They should also be corrosion resistant and strong enough to take
screw attachments.
6. Tool materials should have high wear resistance, good elastic and fatigue strength
properties, and have optimum values of toughness and hardness for the application.
7. The slurry acts as a coolant for the horn, tool and workpiece, supplies fresh abrasive
to the cutting zone and removes debris from the cutting area. It also provides a good
acoustic bond between the tool, abrasive and workpiece, allowing efficient energy trans-
fer.
8. 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.
9. The optimum static load for the maximum machining rate has been found to be depen-
dent on the tool configuration (e.g. cross sectional area and shape), the amplitude and
mean grit size.
Review on ultrasonic machining 253

10. The abrasive material should be harder than the workpiece and, typically, larger abras-
ive grit sizes and higher slurry concentrations yield higher MRR.
11. Low workpiece fracture toughness or increasing the ratio of tool hardness to Young’s
modulus will result in a higher MRR.
12. The superior performance of RUM over standard USM may be explained by the
combined effects of indentation of the workpiece surface, sliding contact between the
embedded grains on the tool/workpiece and rolling contact between the free abras-
ive grains/workpiece.
13. Horn and tool design play an important role in providing a resonant USM system to
maximise the material removal.

Acknowledgements—The authors would like to thank Professors K. B. Haley, Head of the School of Manufactur-
ing and Mechanical Engineering and M. H. Loretto, Director of the IRC in Materials for High Performance
Applications for the provision of laboratory facilities. Additional thanks go to Dr. J. Woodthorpe at T and N
Technology Limited and A. Corfe and D. Jones at Rolls-Royce plc for their technical advice and financial
support. Finally thanks go to the Committee of Vice-Chancellors and Principals (CVCP) of the Universities of
the United Kingdom for the award of an Overseas Research Scholarship (ORS) and the School of Manufacturing
and Mechanical for the provision of a maintenance grant.

REFERENCES
[1] Moreland, M. A., Versatile performance of ultrasonic machining. Cer. Bull., 1988, 67(6), 1045–1047.
[2] Gilmore R., Ultrasonic machining and orbital abrasion techniques, SME Technical Paper (Series) AIR,
NM89-419, 1989, pp. 1–20.
[3] Neppiras, E. A. and Foskett, R. D., Ultrasonic machining – II. Operating conditions and performance of
ultrasonic drills. Philips Tech. Rev., 1957, 18(12), 368–379.
[4] Moore D., Ultrasonic impact grinding, Proc. Non-traditional Machining Conf., Cincinnati, 1985, pp.
137–139.
[5] Koval’chenko, M. S., Paustovskii, A. V. and Perevyazko, V. A., Influence of properties of abrasive
materials on the effectiveness of ultrasonic machining of ceramics. Sov. Powder Metallurgy and Metal
Cer., 1986, 25(7), 560–562.
[6] Gilmore R., Ultrasonic machining of ceramics, SME Paper, MS90-346, 1990, 12 p.
[7] Rozenberg L. D., Kazantsev V. F. et al., Ultrasonic cutting, Consultants Bureau, New York, 1964.
[8] Kohls, J. B., Ultrasonic manufacturing process: Ultrasonic machining (USM) and ultrasonic impact grind-
ing (USIG). The Carbide and Tool J., 1984, 16(5), 12–15.
[9] Haslehurst M., Manufacturing technology (3rd edn.), 1981, pp. 270–271.
[10] Soundararajan, V. and Radhakrishnan, V., An experimental investigation on the basic mechanisms
involved in ultrasonic machining. Int. J. MTDR, 1986, 26(3), 307–321.
[11] Satyanarayana, A. and Krishna Reddy, B. G., Design of velocity transformers for ultrasonic machining.
Electrical India, 1984, 24(14), 11–20.
[12] Drozda T. J. and Wick C., Non-traditional machining – Book chapter 29: Tool and manufacturing engin-
eers handbook (Desk Ed.), Soc. Manuf Engrs, Dearborn, MI, ISBN No. 0872633519, 1, 1983, pp. 1–23.
[13] Pentland, E. W. and Ektermanis, J. A., Improving ultrasonic machining rates – some feasibility studies.
J. Engg. for Ind., Trans. of the ASME, 1965, 87(Series B), 39–46.
[14] Lehman R. L., Primer on engineering ceramics, Ad. Materials and Processes, 141(6), 1992.
[15] Brook R. J., Concise encyclopedia of advanced ceramic materials, Pergamon Press, 1991, pp. 1–2 and
p. 488.
[16] Kennedy, D. C. and Grieve, R. J., Ultrasonic machining – A review. The Prod. Engineer, 1975, 54(9),
481–486.
[17] Perkins J., An outline of power ultrasonics, Technical Report by Kerry Ultrasonics, 1972, 7 p.
[18] Farago, F. T., Abrasive methods engineering. Industrial Press, 1980, 2, 480–481.
[19] Moreland M. A., Ultrasonic impact grinding: What it is: What it will do, 1984 Proc. – 22nd Abrasive
Engg. Soc. Conf.: Abrasives and Hi-Technology, A 2-way Street, 1984, pp. 111–117.
[20] Seah, K. H. W., Wong, Y. S. and Lee, L. C., Design of tool holders for ultrasonic machining using FEM.
J. Materials Processing Technology, 1993, 37(1-4), 801–816.
[21] Markov, A. I., Kinematics of the dimensional ultrasonic machining method. Machines and Tooling, 1959,
30(10), 28–31.
[22] Kainth, G. S., Nandy, Amitav and Singh, Kuldeep, On the mechanics of material removal in ultrasonic
machining. Int. J. MTDR, Pergamon Press, 1979, 19, 33–41.
[23] Kremer, D., The state of the art of ultrasonic machining. Annals of the CIRP, 1981, 30(1), 107–110.
[24] Gilmore, R., Ultrasonic machining: A case study. 7 Int Conf on Computer-Aided Prod. Engg., 1991, 28(1-
2), 139–148.
[25] Moreland M. A., Ultrasonic machining – Book chapter: Ceramics and glasses, ASM International, Engin-
eering Material Handbook, 4, pp. 359–362, ISBN 0871702827, Schneider and Samuel J. (1991).
[26] Nishimura, G., Ultrasonic machining – Part I. J. Fac. Engg. Tokyo Univ., 1954, 24(3), 65–100.
[27] Neppiras E. A., Report on ultrasonic machining, Metalworking Prod., 100, pp. 1283–1288, 1333–1336,
1377–1382, 1420–1424, 1464–1468, 1554–1560, 1599–1604 (1956).
[28] Weller, E. J., Non-traditional machining processes (2nd edn.), Soc. of Manuf. Engineers, 1984, pp. 15–71.
[29] Farrer, J. O., English Patent No. 602801 from 3.6.1948 – USM.
254 T. B. Thoe et al.

[30] Scab, K. H. W. et al., Parametric studies of ultrasonic machining, SME Tech. paper, MR90-294, 1990,
11 p.
[31] Neppiras, E. A., Macrosonics in industry: 1. Introduction. Ultrasonics, 1972, 10, 9–13.
[32] Balamuth, L., Ultrasonic vibrations assist cutting tools. Metalworking Prod., 1964, 108(24), 75–77.
[33] Kremer, D., New developments on ultrasonic machining, SME Technical Paper, MR91-522, 1991, 13 p.
[34] Clifton, D., Imal, Y. and McGeough, J. A., Some ultrasonic effects on machining materials encountered
in the offshore industries, Proc. 30th Int. MATADOR Conf., 1993, pp. 119–123.
[35] Shaw, M. C., Ultrasonic grinding. Microtechnic, 1956, 10(6), 257–265.
[36] Neppiras, E. A., Ultrasonic machining and forming. Ultrasonics, 1964, 2, 167–173.
[37] Machining data handbook (3rd edn.), Compiled by the Technical Staff of the Machinability Data Centre,
Cincinnati Metcut Research Associates Inc., 1980, 2, 43–63.
[38] Venkatesh, V. C., Machining of glass by impact processes. J. Mech. Working Technology, 1983, 8,
247–260.
[39] Hartley, M. S., Ultrasonic machining of brittle materials. Electronics, 1956, 29(1), 132–135.
[40] McGeough, J. A., Advanced methods of machining, Chapman and Hall, ISBN 0412319705, 1988, pp.
170–198.
[41] Komaraiah, M., Manan, M. A., Narasimha Reddy, P. and Victor, S., Investigation of surface roughness
and accuracy in ultrasonic machining. Precision Engg., 1988, 10(2), 59–65.
[42] Balamuth, L. A., Ultrasonic assistance to conventional metal removal. Ultrasonics, 1966, 4, 125–130.
[43] Graff, K. F., Macrosonics in industry: 5. ultrasonic machining. Ultrasonics, 1975, 13, 103–109.
[44] Ultrasonics – can aid conventional machining, Metalworking Prod., 1962, 106(38), 55-56.
[45] Isaev, A. I., Learning with ultrasonically vibrated reamers. Machines and Tooling, 1962, 33(6), 27–30.
[46] Moreland, M. A., Ultrasonic advantages revealed in the hole story. Cer. Appl. Manuf., 1988, 187, 156–162.
[47] Murti, V. S. R. and Philip, P. K., A comparative evaluation of ultrasonic assisted EDM through topograph-
ical indices, 12th AIMTDR Conf., IIT Delhi, Tata McGraw Hill, New Delhi, 1986, pp. 365–369.
[48] Halm, R. and Schulz, P., Ultrasonic machining of complex ceramic components, Erosion AG Report,
DKG 70. No. 7, 1993, 6p.
[49] Black, P., An ultrasonic impact grinding technique for electrode-forming and redressing, Proc. Non-tra-
ditional Machining Conf., Cincinnati, Ohio, ASM, 1985, pp. 129–136.
[50] Kremer, D., Bazine, G. and Moison, A., Ultrasonic machining improves EDM. technology, Electromachin-
ing, Proc. 7th Int. Symp., ed. by Prof. Crookall J. R., Birmingham, UK, 1983, pp. 67–76.
[51] Ghabrial, S. R., Trends towards improving surfaces produced by modem processes, Wear, 109, No. 1–4,
Paper presented at the 3rd Int. Conf. on Metrol and Prop. of Eng’g Surf, Teesside, Engl., 1986, pp.
113–118.
[52] Thoe, T. B., Ultrasonic contour machining of ceramic materials, MPhil Thesis, Uni. of Birmingham (1994).
[53] Gilmore, R., Ultrasonic machining, SME Technical Paper, EM89-123, 1989, 10p.
[54] Essam, G. R., The application of rotary ultrasonic machining to ceramic turbine blades, Rolls Royce Lab’y
Rep., LLR 30898 (1984).
[55] Iwanek, H., Grathwohl, G., Hamminger, R. and Brugger, N., Machining of ceramics by different methods,
Cer. Mat’ls and Components for Engines, Proc. 2nd Int. Symp., 1986, pp. 417–423.
[56] Oh, H. L., Pankow, D. and Finnie, I., The mechanism of ultrasonic machining of brittle solids, Proc. Int.
Conf. Prod. Eng., Tokyo, Part II, 1974, pp. 109–113.
[57] Jay, F., IEEE standard dictionary of electrical and electronics terms, 3rd edition, 1984, pp. 405 and 519.
[58] Lucas, M., ESPI Analysis of Ultrasonic Machining Sonotrodes – A consultancy report to T and N, Lough-
borough Consultants Limited (1992).
[59] Boothroyd, G. and Knight, W. A., Fundamentals of machining and machine tools (2nd Edition), Marcel
Dekker, 1989, pp. 469–475; 490–499 and 510–515.
[60] Neppiras, E. A. and Foskett, R. D., Ultrasonic machining – I. Technique and equipment. Philips Tech.
Rev., 1957, 18(11), 325–334.
[61] Legge, P., Machining without abrasive slurry. Ultrasonics, 1966, 4, 157–162.
[62] Ed. by L. D. Rozenberg, Physical principles of ultrasonic technology – Vol. 1 and 2, Plenum Press, N.
Y. (1973).
[63] Frederick, J. R., Ultrasonic engineering, John Wiley and Sons Inc., New York, ISBN 0471277258 (1965).
[64] Kaczmarek, I., Impact Grinding (Ultrasonic machining) – Book chapter 21: Principles of machining by
cutting, abrasion and erosion, Peter Peregrinus Ltd., Stevenage, ISBN 0901223662, 1976, pp. 448–462.
[65] Rawson, F. F., High power ultrasonic resonant horns, part 1 – Basic design concepts. velocity of ultrasound
at 20 KHz; effects of material and horn dimensions, Ultrasonics Int. 87 Conf. Proc., London, 1987, pp.
680–685.
[66] Prabhakar, D. and Haselkorn, M., An experimental investigation of material removal rates in rotary ultra-
sonic machining. Trans. of NAMRI/SME, 1992, 20, 211–218.
[67] Amin, S. G., Ahmed, M. H. M. and Youssef, H. A., Optimum design charts of acoustic horns for ultrasonic
machining. Proc. Int. Conf on AMPT’93, 1993, 1, 139–147.
[68] Thoe, T. B., Aspinwall, D. K. and Wise, M. L. H., The effect of operating parameters when ultrasonic
contour machining, Twelfth Annual Conference of the Irish Manufacturing Committee (IMC-12), Cork,
Ireland, September 1995, pp. 305–312.
[69] US400 – Ultrasonic machining system, Brochure from Erosonic AG (1994).
[70] Wojchiechowski, M. P. et al., Ultrasonic machining: Past, present and future, SME Paper, MR72-188,
1972, 12p.
[71] Adithan, M., Tool wear studies in ultrasonic drilling. Wear, 1974, 29, 81–93.
[72] Kumehara, H., Characteristics of threaded joints in ultrasonic vibrating system. Bull. of JSME, 1984,
27(223), 117–123.
[73] Adithan, M., Production accuracy of holes in ultrasonic drilling. Wear, 1976, 40(3), 309–318.
Review on ultrasonic machining 255

[74] Kazantsev, V. F., Improving the output and accuracy of ultrasonic machining. Machines and Tooling,
1966, 37(4), 33–39.
[75] Ultrasonic machining of glass at the N. P.L., Machinery, May 1964, pp. 1172–1176.
[76] Adithan, M., Abrasive wear in ultrasonic drilling. Tribology International, 1983, 16(5), 253–255.
[77] Goetze, D., Effect of vibration amplitude, frequency and composition of the abrasive slurry on the rate
of ultrasonic machining in Ketos tool steel. The J. Acoust. Soc. of Am., 1956, 28(6), 1033–1037.
[78] Krar S. F. and Ratterman E., Superabrasives – Grinding and machining, McGraw Hill, 1990, pp. 12.
[79] Miller, G. E., Special theory of ultrasonic machining. J. Appl. Phy., 1957, 28(2), 149–156.
[80] Cook N. H., Manufacturing analysis, Addison–Wesley, New York, 1966, pp. 133–138.
[81] Khairy, A. B. E., Assessment of some dynamic parameters for the ultrasonic machining process. Wear,
1990, 137, 187–198.
[82] Ghabriel, S. R., Saleh, S. M., Kohail, A. and Moisan, A., Problems associated with electrodischarge
machined electrechemically machined and ultrasonically machined surfaces. Wear, 1982, 83, 275–283.
[83] Komaraiah, M and Narasimha Reddy, P, A study on the influence of workpiece properties in ultrasonic
machining. Int. J. Mach. Tools Manuf, 1993, 33(3), 495–505.
[84] Enomoto, Y., Sliding fracture of soda–lime glass in liquid environments. J. Mat’l Science, 1981, 16(12),
3365–3370.
[85] Lucas M. and Chapman G. M., Vibration analysis at ultrasonic frequencies, The 1989 ASME Ds. Tech’l
Conf.–12th Biennial Conf on Mech’l Vib. and Noise,Montreal, DE-Vol. 18–4, September 1989, pp.
235–240.
[86] Smith, T. J., Parameter influence in ultrasonic machining. Ultrasonics, 1973, 11(5), 196–198.
[87] Markov A. l., Ultrasonic machining of intractable materials Iliffe Books Ltd., London (1966).
[88] Adithan, M. and Venkatesh, V. C., Study of the performance characteristics of an ultrasonic drilling head.
Wear, 1975, 33, 261–270.
[89] Kubota, M., Tamura, Y. and Shimamura, N., Ultrasonic machining with a diamond impregnated tool.
Bull. Japan. Soc. of Prec. Engg., 1977, 11(3), 127–132.
[90] Komaraiah, M. and Reddy, P., Narashima, Rotary ultrasonic machining – A new cutting process and its
performance. Int. J. Prod. Res., 1991, 29(11), 2177–2187.
[91] Takeyama, H., Burrless drilling by means of ultrasonic vibration. Annals of CIRP, 1991, 40(1), 83–86.
[92] Kops, L., Investigation into the influence of the wear of the abrasive powder on the technological indices
of ultrasonic machining. CIRP Annalen, Springer Verlag, Berlin, 1964, 12(3), 151–157.
[93] Kazantsev, V. F., The relationship between output and machining conditions in ultrasonic machining.
Machines and Tooling, 1963, 34, 14–17.
[94] Smith, T. L., Parameter influence in ultrasonic machining, BSc(Hon) Diss., The Nottingham Trent Uni.,
(1971),
[95] Ghabrial, S. R., Saleh, S. M., Moisan, A. and Kremer D., Some aids towards improving performance in
U. S.M. 1984 SME Manuf Engg. Trans., 12th NAMRC Conf. Proc., 1984, pp. 227–232.
[96] Markov, A. I., Ultrasonic drilling and milling of hard non-metallic materials with diamond tools. Machines
and Tooling, 1977, 48(9), 45–47.
[97] Komaraiah, M. and Reddy, P. N., Relative performance of tool materials in ultrasonic machining. Wear,
1993, 161(1-2), 1–10.
[98] Adithan, M., Tool wear characteristics in ultrasonic drilling. Tribology Int., 1981, 14(6), 351–356.
[99] Adithan, M. and Venkatesh, V. C., Parameter influence on tool wear in ultrasonic drilling. Tribology Int.,
1974, 7(6), 260–264.
[100] Riddei, V., Cavitation erosion – A survey of the literature, 1940–1970. Wear, 1973, 23, 133–136.
[101] Nandi, G., Mukherjee, S. K. et al., Tool wear in ultrasonic trepanning of glass-ceramics, The Japan soc.
of Precision Engg., Proc. 5th Int. Conf on Prod. Engg. Tokyo 1984, 1984, pp. 283–287.
[102] Clouser, H. A., New techniques for electrode forming, Res. and Technological Dev. in Non-traditional
Machining Presented at the Winter Annual Meeting of the ASME, 1988, pp. 75–88.
[103] Dam, H. et.al., Surface characterization of ultrasonic machined ceramics with diamond impregnated sono-
trode, Proc. the Int. Conf. on Machining of Adv. Mat’ls, Gaithersburg, Maryland, July 1993, pp. 125–133.
[104] Ultrasonic assembly of thermoplastic mouldings and semi-finished product – Recommendations on
methods, construction and applications. Manual written by German Electrical Manufacturers Associ-
ation ZVEI.
[105] Merkulov, L. G., Design of ultrasonic concentrations. Akusticheskiy Zhurnal, 1957, 3, 246–255.
[106] Pandey, P. C. and Shan, H. S., Modern machining processes, Tata McGraw-Hill, Chapter 2, 1980, pp.
7–38.
[107] Amin, S. G., Ahmed, M. H. M. and Youssef, H. A., Computer aided design of acoustic horns for ultrasonic
machining using finite element analysis. Proc. AMPT93 Conf., 1993, 2, 1455–1465.
[108] Hahn, R., Ultrasonic machining of glass and ceramics. Am. Cer. Soc. Bull., 1993, 72(8), 103–106.

You might also like