You are on page 1of 33

Handbook of Manufacturing Engineering and Technology

DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Loose Abrasive Machining


Takashi Satoa,b*, Swee Hock Yeob and Hamid Zarepourb
a
Singapore Institute of Manufacturing Technology (SIMTech), Singapore, Singapore
b
School of Mechanical and Aerospace Engineering, Nanyang Technological University (NTU), Singapore, Singapore

Abstract
Loose abrasive machining is one of the processes, which contributes to improving precision, such as
surface roughness and form accuracy of manufactured components. To date, numerous process
principles have been developed to materialize the loose abrasive machining process in different
ways. These developments have been made in response to the changing needs of the commercial
market and to the improvement of the quality of products. This chapter provides introductions of
processes for ultrasmooth surface, complex geometry, and mass finishing as well as the latest
discussions and findings about ultrasonic machining.

Introduction
Many types of machining process have been developed and utilized by researchers and engineers
with objectives to improve devices’ performance, function, economic, compact, light-weight, and
design.
Loose abrasive machining is one of the processes, which contributes to improving precision, such
as surface roughness and form accuracy of manufactured components.
Loose abrasive process is categorized as a constant pressure process. Since abrasive is not fixed,
material removal depends on the amount of pressure that is set. On the other hand, fixed abrasive
process such as grinding is categorized as a constant depth-of-cut process. The amount of material
removal will therefore depend on how much the depth of cut is set.
Preston’s equation is a well-known formula in explaining the relationship between material
removal rate (MRR) during process relative to main parameters (Preston 1927). This is useful in
the understanding of material removal mechanism under constant pressure process. According to
Preston’s equation, the MRR is proportional to the product of the polishing pressure, p, and relative
velocity between workpiece and polishing tool, v:

dh
MR ¼ ¼kpv
dt
where h is the amount of material removal; t is the processing time; and k is the Preston coefficient.

*Email: sato@simtech.a-star.edu.sg
*Email: taka2889@gmail.com

Page 1 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 1 Schematic diagram of lapping and polishing (a) Lapping (b) Polishing

Since this equation is adaptable to most of the loose abrasive processes, it is useful to aid in the
understanding and estimation of the processed results.
In this chapter, processes for ultrasmooth surface, p complex geometry, and mass finishing are
introduced respectively. Followed by a machining process, using ultrasonic is also introduced with
detail information.

Process for an Ultrasmooth Surface

Lapping and Polishing


Lapping and polishing are loose abrasive finishing processes, which are among the commonly used
methods for fabrication of extremely smooth surfaces, e.g., a critical requirement of optical lenses,
bearings, dies and molds, silicon wafers, and other precision components (Nakazawa 2004;
Venkatesh and Izman 2008). Figure 1 shows the schematic diagram of lapping and polishing.
Lapping is applied to produce a smooth surface. The lapping uses a tool, called a lap, which has
a reverse shape of the desired form. The lap rubs the loose abrasive which is generally suspended in
fluid, against the workpiece surface, and replicates the tool form to the workpiece while smoothen-
ing the surface. In general, surface with Ra ¼ 0.04 to 0.01 mm are obtained as a result of lapping.
Laps are generally made of cast iron, soft steel, brass, hardened steels, as well as glass. Copper and
steel laps enable the lapping process to be accelerated, but cast iron laps retain their shape better and
produce a smoother surface at the same time. Glass laps are capable to remove high levels of metal
when used with fine-grain abrasives and produce an even better surface than cast iron laps.
Lapping media that are commonly used are aluminum oxide, silicon carbide, emery, boron
carbide, and pastes or compounds mixed with oil or bonding vehicle (gasoline, kerosene, vegetable
oil, etc.)
Polishing compared to lapping is applied to produce extremely smooth surface. The polishing
process uses a tool, relatively soft tools, such as pads made of soft cloth or resin which is used to rub
loose abrasive slurry against the workpiece surface thus improve the smoothness of workpiece.

Chemical Mechanical Polishing (CMP)


Chemical mechanical polishing (CMP) was introduced by IBM for the production of the 64 Mbit
DRAM chip (Zantyea et al. 2004). This process is frequently used for the planarization of

Page 2 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 2 Schematic diagram of CMP

semiconductor wafers; CMP takes advantage of the synergetic effect of both mechanical and
chemical forces in the polishing of wafers (Patrick et al. 1991; Zantyea et al. 2004; Matijevic and
Babu 2008; Venkatesh et al. 1995). This is performed by applying a load force to the back of a wafer
while it rests on a pad. Both the pad and wafer are then counter rotated while introducing slurry
containing both abrasives and reactive chemicals underneath.
A schematic diagram of CMP is shown in Fig. 2. A wafer containing the films, which is
a workpiece, is mounted in a carrier and pressed facedown at a known pressure and rotated against
a porous polyurethane pad mounted on a rotating table. Both the carrier and the platen are normally
rotated in the same direction. When the rotational speed of the carrier and the platen are the same, the
relative velocity of each point on the wafer with respect to the pad is the same, facilitating a uniform
material removal from across the entire wafer surface. After determining that the proper amount of
material has been removed by using one of a variety of end point determination techniques, the wafer
is removed from the carrier and washed.
To obtain the better quality surface, this set of CMP process and cleaning processes are repeated
for each level of metallization. It is useful to remember that in a typical CMP process, the thickness
of the material removed is limited to about a micrometer, with a removal uniformity of perhaps a few
tenths of a nanometer across a heterogeneous surface on a 300 mm diameter silicon wafer – a daunting
and inherently complex task.
A large number of parameters influence the outcome of the CMP process. These can be divided
into two groups. The first group, which depends on the polishing tool configuration, includes the
design of the wafer carrier and retainer ring, applied pressure and its distribution across the wafer,
rotational speeds of the carrier and the platen, etc. The second group consists of those dictated by the
consumables used during the process, which include the characteristics of the pad like its hardness,
modulus, porosity, surface roughness, grove design, conditioning, etc., and the large number of
properties associated with the slurry. The latter contains abrasives (e.g., silica, alumina, ceria,
zirconia, etc.) as well as several different chemical additives. The components of the slurry, i.e.,
the abrasive size, shape, method of preparation and concentration, as well as the chemical additives,
and more importantly the pH strongly influence the polishing process in terms of both removal rates
and defects caused.

Elastic Emission Machining (EEM)


EEM is one of the atomic size machining methods (Mori et al. 1987; Mori et al. 1988). When two
solid phase materials composed of different chemical elements come into contact with each other,
different kinds of interactions are being generated at the interface. Therefore, if these solids are

Page 3 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 3 Schematic diagram of EEM

separated mechanically, chances that the atoms from one surface moving onto the other surface
might occur. This kind of phenomenon is applied in machining and is known as EEM.
Ultrafine powder particles consisting of diameters that are much smaller than 1.0 mm are
homogeneously mixed with water. Making use of the flow of this mixture, powder particles are
accelerated and transported onto the work surface with minimal load (Fig. 3). When in contact with
the work surface, surface atoms will be removed through the process mentioned above. Limitation of
working area is within the contacting area which needs to be smaller than 10 nm2, and removal is
possible only where mutual surface atoms are ideally binding.
Furthermore, to remove the target atom, the interface has to possess the characteristics to decrease
the binding energy between the atoms in the surface and second layers. As a result, to obtain
a geometrically perfect surface, both machines area and the depth need to be of approximate atomic
order. The interface characteristics have shown that the removal of atoms from the work surface is
semi-spontaneous, thus finished surfaces from the point of view of physical properties can be
perfect.

Process for Complex Geometries

Limitation of the flexibility and motion of the tool configuration make lapping and polishing not
practical for the treatment of complex surfaces, such as sharp corners, deep recesses, sharp pro-
jections, free-form surfaces, and interiors of complicated components.
Components used in critical applications in the aerospace, biomedical, and semiconductor
industries require highly finished surfaces to achieve their desired surface functions, and many of
these are currently accomplished manually despite increasing production costs. In fact, some of
these are virtually unreachable by conventional techniques, and the lack of finishing technology acts
as an obstacle to the technology innovation.

Abrasive Flow Machining (AFM)


AFM Process
Abrasive flow machining (AFM) was developed by Extrude Hone Corporation, USA, in 1960.
Three types of AFM machines that have been reported in the literature are one-way AFM, two-way
AFM, and orbital AFM. The most common used AFM is two-way AFM in which two vertically

Page 4 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 4 Schematic diagram of AFM

opposed cylinders extrude medium back and forth through passages formed by the workpiece and
tooling (Fig. 4).
AFM is used to deburr, radius, and polish difficult to reach surfaces by the extrusion of an
abrasive-laden polymer medium with very special rheological properties through the workpiece
repeatedly. It is widely used in finishing processes to finish intricate shapes and profiles (Loveless
et al. 1994; Raju et al. 2005). The polymer abrasive medium used in this process possesses easy
flowability, good self-deformability, and fine abrading capability. Layer thickness of the material
removed is of the order of about 1–10 mm. The best surface finish that has been achieved is 50 nm
and tolerances are  0.5 mm. In this process, tooling plays a tremendous role in the finishing of
material; however, hardly any literature is available on this kind of process. In AFM, deburring,
radiusing, and polishing are performed simultaneously in a single operation on various areas
including conventionally inaccessible areas, and it can produce true round radii even on complex
edges.
AFM reduces surface roughness by 75–90 % on cast and machined surfaces. It can process dozens
of holes or multiple passage parts simultaneously and achieve uniform results. For example, air
cooling holes on a turbine disk and hundreds of holes in a combustion liner can be deburred and
radiused in a single operation.
AFM maintains flexibility while capable to perform jobs which require hours of highly skilled
hand polishing in a few minutes; AFM produces uniform, repeatable, and predictable results on an
impressive range of finishing operations. An important feature which differentiates AFM from other
finishing processes is that it is possible to control and define the intensity and location of abrasion
through fixture design, medium selection, and process parameters. It has wide applications in many
areas such as aerospace, dies and moulds, and automotive industries.

Magneto-rheological Abrasive Flow Finishing (MRAFF)


Magneto-rheological abrasive flow finishing (MRAFF) was developed by Jha et al. in 2004 as a new
precision finishing process using MR fluid for complicated geometries (Jha et al. 2007). This process
employs determinism and in-process controllability of the rheological behavior of an abrasive-laden

Page 5 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 5 Schematic diagram of the MRAFF

medium to finish intricate shapes. The desired properties of the base MR fluid used in MRAFF are as
follows:

1. The fluid should be thermally stable and should have a high boiling point
2. It should be noncorrosive and nonreactive with the employed magnetic and abrasive particles

The representative MR fluid slurry consists of carbonyl iron particles (6 mm in mean diameter:
20 vol.%), silicon carbide abrasive (mesh size #800, #1,000, #1,200, or #1,500: 20 vol.%), and
organic medium (60 vol.%).
In the MRAFF process, magnetically stiffened slug of MR polishing fluid is being extruded back
and forth with a piston through or across the passage formed by the workpiece and fixture. Selective
abrasion occurs only where the magnetic field is applied across the workpiece surface while keeping
the other areas unaffected. The schematic diagram of the process is shown in Fig. 5. Looking into the
rheological behavior of the polishing fluid, it changes from nearly Newtonian to Bingham plastic
and back when entering, traversing, and exiting the finishing zone, respectively. The abrasive cutting
edges, which are held by carbonyl iron chains, rub the workpiece and shear the peaks away from its
surface. The bonding strength of the field-induced structure of the MR polishing fluid and the
extrusion pressure applied through the piston determine the amount of material sheared from the
workpiece surface peaks by the abrasive grains.

Electromagnetic Field-Assisted Machining


Magnetic Field-Assisted Finishing (MFAF)
The combination of a magnetic field with the mechanical action of a magnetic tool against
a workpiece gives rise to the magnetic field-assisted finishing (MAF) process (Ko et al. 2003; Yin
and Shinmura 2004; Yamaguchi et al. 2007). The magnetic tools can be introduced into areas that are
hard to reach by conventional technologies and by means of magnetic manipulation; they exhibit
relative motion against the workpiece surface needed for finishing. This shows potential for
overcoming problems associated with more conventional finishing processes.
Figure 6 shows the schematic diagram of MAF. The magnetic abrasive (magnetic tools) consists
of iron particles and Al2O3 abrasive grains. The composite ingot is produced from the thermite
process using aluminum powder and iron oxide powder. Subsequently, the ingot is then mechani-
cally crushed and sieved to form the finished magnetic abrasive. The Al2O3 grains are contained
both inside and outside of the resulting magnetic abrasive.

Page 6 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 6 Schematic diagram of MAF (a) Utility of magnetic abrasive blush. (b) Utility of magnetic transmission
phenomenon

In a magnetic field, ferrous particles (including magnetic abrasive) suspended by magnetic force
are linked together along the lines of magnetic flux. When the magnetic flux flows unimpeded
through the nonferrous workpiece material, it might influence the motion of a ferrous particle – even
if the particle is not in direct contact with a magnetic pole – and the magnetic field can be controlled
from outside. The ferrous particle chains connected by magnetic force allows a flexible configura-
tion, and given this unique behavior of the ferrous particles, it enables the application of the finishing
operation to easily accessible surfaces and also to areas that are hard to reach by means of
conventional mechanical techniques.

Float Polishing Using Magnetic Fluid


Magnetic fluids (MFs) were developed by Papell in 1965 to magnetically control fuel flow for the
Apollo project in the zero gravity conditions of space (Papell 1965). They are found in applications
such as seal components and are routinely used in voice coils, dampers, and rotary brakes.
MFs are stable colloidal suspensions of permanently magnetized particles, such as magnetite.
This stability comes from the Brownian motion which keeps the particles, which are about 10 nm in
diameter, from settling under gravity, and at the same time, a surfactant covers each particle to create
short-range steric repulsion between particles, which prevents particle agglomeration in the presence
of nonuniform magnetic fields.
Application of a MF to polishing and finishing processes was undertaken by Kurobe and Inamaka
in 1983 (Kurobe et al. 1983). The resultant process, called magnetic field-assisted fine finishing, was
developed as a new lapping technique for the controllable finishing of materials, such as semi-
conductors and ceramics in the electronics and precision machinery fields. Figure 7 shows the
schematic diagram of magnetic float polishing. The MF engulfs the groove cut in the brass disk, and
the polisher (a 1 mm thick rubber sheet) covers the magnetic fluid-filled groove. After which, the
water-based polishing compound was supplied over the polisher. When DC voltage is applied,
electromagnets placed above and below the disk create a magnetic field. As shown in Fig. 7, the
polishing pressure is generated by the ferrous particles, which are attracted by the magnetic field, and
pushes the polishing compound (via the polisher) against the work surface. The polishing action
results when the upper pole, which is connected to the workpiece, and the disk rotate in opposite
directions. This method has been experimentally proven that flexible surface finishing is possible,
and the surface roughness and stock removal rate can be controlled by changing the current to the
electromagnet.

Page 7 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 7 Schematic diagram of magnetic float polishing

Fig. 8 Schematic diagram of the MRF processing principle

Magneto-rheological Finishing (MRF)


The initial discovery and development of magneto-rheological fluids (MR fluids) and devices is
credited to Jacob Rabinow, who studied them at the US National Bureau of Standards (now named
the National Institute of Standards and Technology (NIST)) in the 1940s (Rabinow 1948). MR fluids
typically consist of micron-sized, magnetically polarizable particles dispersed in a carrier medium;
some common media are mineral oil or silicone oil. Particle chains form when a magnetic field is
applied. As a result, the viscosity of the MR fluids apparently increases (Kurobe et al. 1983).
The MRF process was firstly introduced commercially in 1998, and the manufacture of precision
optics has since changed dramatically (Tricard et al. 2003). The developed process makes use of
polishing slurry based on MR fluid, which can be mixed, pumped, and conditioned in their liquid
state; however, in the presence of an applied magnetic field, it causes a viscosity change to
a semisolid state creating a stable and conformable polishing tool. A typical composition of an
MR fluid is 36 % carbonyl iron, 6 % abrasive (cerium oxide), 3 % stabilizer, and 55 % water.
Figure 8 shows the schematic diagram of the MRF processing principle. A workpiece is fixed at
some distance from a moving surface, in order to ensure that the workpiece surface and the moving
surface form a converging gap. In the area around the gap, a nonuniform magnetic field is generated

Page 8 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 9 Schematic diagram of processing principle of internal finishing using MRF-based slurry

when an electromagnet is placed below the moving surface. The MR fluid is delivered to the moving
surface just above the electromagnetic poles and then pressed against the surface by the magnetic
field gradient making the fluid a Bingham plastic before it enters the gap. Thereafter, the shear flow
of plastic MR fluid flows through the gap, resulting in the development of high stresses in the
interface zone and thus, as a result, material removal occurs over a portion of the workpiece surface.
The process can make significant improvements to the surface roughness, and flatness has since
been adopted by major manufacturers of precision optics. In the aspect of both form accuracy and
micro-roughness, the MRF process has demonstrated the ability to produce optical surfaces to tight
tolerances. The surfaces, including aspheres, can be made with materials ranging from glass/glass
ceramics (including fused silica, ULE, and Zerodur) to single-crystalline materials (including
silicone and calcium fluoride) or polycrystalline materials (including SiC).

Internal Surface Finishing Using Magneto-rheological Fluid-Based Slurry


A new type of slurry – referred to as MRF-based slurry – was developed in 2006 by Yamaguchi and
Sato (Sato et al. 2007) and is especially appropriate for the internal finishing process of piping
systems in micro- and nanotechnological industrial devices. The major feature of the MRF-based
slurry is that the abrasives are smaller than the iron particles so that the iron particles can trap as
many abrasive particles as possible.
The processing principle is illustrated in Fig. 9. The MRF-based slurry is introduced inside the
work and is attracted to the finishing area by the field generated by magnetic poles (e.g., electro-
magnetic coils or permanent magnets). When the work is rotated and oscillated along its axis, there is
relative motion between the abrasives and the work surface, and the entire inner workpiece surface
can be finished. Under typical finishing conditions, the MRF-based slurry experiences a magnetic
force, a centrifugal force, a friction force against the work surface, and the force of gravity; these
combine and result in the dynamic behavior shown in Fig. 9b. This behavior continuously displaces
the cutting edges of the abrasives in the MRF-based slurry, which improves the finishing efficiency.

Page 9 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 10 Schematic diagram of MCF float polishing

Magnetic Compound Fluid Finishing (MCF)


Magnetic compound fluid (MCF), which basically consists of MF and MR fluid, was developed by
Shimada et al. in 2001 as an intelligent fluid (Shimada et al. 2008). In the same conditions, the
apparent viscosity of MF in a magnetic field is lower than that of MR fluid. However, the stability of
the particle distribution in MF is better than in MR fluid. Therefore, by changing the mixing ratio of
MF and MR fluid, the apparent viscosity and the particle distribution stability in MCF can be altered.
This characteristic must be the greatest advantage of MCF. The structure of MCF is made up of
chain-shaped magnetic clusters consisting of magnetic particles of different sizes. Magnetic particles
of MF (a few nm in diameter) surround the magnetic particles of MR fluid (a few mm in diameter),
and long clusters of magnetic particles are formed in the process.
Research into the application of MCF to float polishing has been conducted from 2002 (Shimada
et al. 2008; Sato et al. 2010). Figure 10 illustrates schematic diagram of MCF float polishing and
chain-shaped clusters consisting of MCF mixed with abrasive. As the MCF is subjected to
a magnetic field, chain-shaped magnetic clusters composed of nanometer-sized magnetic particles
and micrometer-sized magnetic particles are thus created along the lines of magnetic flux. The
dimensions of the clusters depend on the composition of the MCF, the strength of the magnetic field,
and the method by which the field is applied. The nonmagnetic abrasive particles are trapped by the
clusters or distributed between clusters, and to increase the viscosity, alpha cellulose fiber has been
used and interspersed with the clusters. When the magnet is moved, the clusters and abrasive exhibit
relative motion against the work surface, and material removal by the micro-cutting action of the
abrasive particles can be observed. This process can be diversely applied to free-form metal surfaces,
ceramics, glasses, resin surfaces, inner surfaces of tubes, etc.

Process for Mass Finishing


Deburring and surface finishing are necessary processes for almost all metal and many plastic parts.
Mass finishing is the term most commonly associated with abrasive tumbling operation. Typically,
large number of parts are placed in the deburring machine, and they are deburred together at the same
time. Equipment operation can be performed as a batch process or a continuous flow process with
the use of several different styles of machines.

Page 10 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 11 Schematic diagram of barrel finishing

Barrel Finishing
Conventional rotary barrel tumbling is the original mass finishing technique. Ancient Chinese and
Egyptians used tumbling barrels with natural stones as media to achieve smooth finishes on weapons
and jewelry (Gillespie 2006). Figure 11 shows the schematic diagram of barrel finishing.
Barrel finishing is now a vastly improved process compared to the ancient tumbling operations,
but this process is slow in general. Given that barrel finishing is a versatile means of edge and surface
conditioning, equipment costs are cheaper and operation is simple; therefore, there are still appli-
cations as this is the most economical process.
Barrel finishing is a low-pressure abrading process generally performed by the controlled sliding
and rolling action of workpieces, media, and compounds. In a rotary or tumbling barrel, the upper
layer of the workload has a sliding movement. While the barrel rotates, the load moves upward in the
barrel to a turnover point. The force of gravity overcomes the tendency of the mass to stick together,
and then the upper layer slides toward the bottom of the barrel. The barrel is normally loaded to
approximately 60 % of capacity with a mixture of workpieces, media, and compounds.
Higher load levels will be better for some workpiece such as large or heavy components. If the
load level is increased above 60 %, this decreases the length of slide, thus the probability of the
workpieces contacting each other will be reduced. However, while the force of workpiece contact is
being reduced, cycles will have to be increased.
Increase in barrel rotation increases the steepness of angle of slide and thus accelerates the action.
However, faster barrel rotation increases the tumbling action, which in turn increases the likelihood
of damaging the workpieces and reducing the quality of edge and surface conditions.

Vibratory Finishing
Vibratory finishing is now the most favored type of mass finishing, right next to hand deburring, the
most common surface conditioning method used by industries (Gillespie 2006). This versatile
process is commonly used for cleaning, deburring, deflashing, descaling, edge and corner radiusing,
surface finishing, and stress relieving.
Figure 12 shows the schematic diagram of vibratory finishing. The first tub-type vibratory
finishing machine was introduced commercially in 1957, and the bowl-type about 5 years later.
Workpieces are loaded into the open top of a container holding the media, compound, and water.

Page 11 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 12 Schematic diagram of vibratory finishing

Fig. 13 Schematic diagram of drag finishing

Vibratory finishing has been widely employed for final surface finishing of products, given credit
to the capability to finish with consistency and with considerably lower manufacturing cost.
Workpieces of wide variety of sizes and shapes are processed and are applicable to all metals and
many nonmetallic materials. The advantage of conducting the process in large quantities in batch or
continuous process setups without handling or fixturing helps to minimize costs.

Drag Finishing
Drag finishing is the process of dragging parts through a bed of media (Gillespie 2006). The parts
can be handled in single or fixture in groups. The motion of the parts through the media can be in
a straight line, an oval-shaped racetrack system, or some planetary motion. Figure 13 shows
a schematic diagram of drag finishing.

Page 12 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 14 Basic elements of the USM system: 1 transducer, 2 booster, 3 horn, 4 tool, 5 workpiece, 6 fixture, 7 slurry pump,
8 slurry tank, 9 slurry nozzle (Weller 1984)

The processes have two variations: first, the parts are fixed on nonrotating fixture and dragged
through the media, and second, the parts are placed on spindles that rotate while being dragged
through the media. The first approach is used in Japan, while the second approach is a common
design in the USA.
The advantage of the process is the non-impingement of the parts on each other and large parts can
be processed. In drag finishing, parts are inserted into a tank of moving abrasive media and
compound and moved in the opposite direction of the media. Similarly as other loose abrasive
processes, the abrasive media and compound rub against the parts and its edges.
Heavy burrs can be removed, however, only with noticeable stock loss and long cycle time. This
process does not work well on edges with deep internal features and impregnates minute particles of
abrasive into the work surface which can result in poor brazed, soldered, or welded joints and
increase the probability of plating failures.

Ultrasonic Machining

Overview of the Ultrasonic Machining


The potential of the sound waves with high frequency (about 70 kHz) for machining was first
observed by R. W. Wood and A. L. Loomis in 1927 (McGeough 1988). Subsequently, the first patent
on using ultrasound for machining was granted to L. Balamuth in 1945 (Balamuth 1945). In the
investigation on the ultrasonic grinding of abrasive powders, the surface of a container filled with
abrasive slurry was found to disintegrate in the areas close to the vibrating tip of an ultrasonic
transducer. In addition, the shape of the generated cavity was exactly the same as that of the
transducer tip (McGeough 1988). Consequently, the capabilities of this technique for machining

Page 13 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 15 Working principle of USM (Zhang et al. 2005; Rajurkar et al. 2006)

applications were quickly recognized by industrial users. The production of ultrasonic machine tools
then started in the early 1950s by mounting the first USM tools on the bodies of drilling and milling
machines. By 1960, various USM tools were commercialized and utilized by manufacturers in
regular production. Various terms have been coined for the USM such as ultrasonic drilling, slurry
drilling, and ultrasonic abrasive machining. However, it is more commonly referred as ultrasonic
impact grinding (USIG), ultrasonic-assisted lapping (USAL), and ultrasonic machining (USM).

Basic Elements and Working Principles of USM


Figure 14 shows the basic elements of an ultrasonic machining system. High-frequency electrical
signal produced by an ultrasonic power generator is converted into mechanical vibration via
a transducer-booster combination. The booster may increase, decrease, or retain the vibration
amplitude received from the transducer. The mechanical vibration is then transmitted to a horn-
tool assembly (known as sonotrode) which acts as an energy-focusing device. This results in
vibration of the tool in the axial direction, typically at a frequency and amplitude ranging from
20 to 40 kHz and 5 to 50 mm, respectively. The power employed in the USM process usually ranges
from 50 to 3,000 W, and it may reach up to 4 kW in some USM systems (Thoe et al. 1998).
Figure 15 illustrates the working principle of the USM process. As illustrated, abrasive slurry
which consists of a powder dispersed in a liquid medium (water or oil) is fed into the machining zone
between the tool and workpiece. Abrasive materials such as boron carbide, boron nitride, aluminum
oxide, silicon carbide, and polycrystalline diamond with particle size of 50–300 mm are commonly
used. A downward controlled force (also known as static load or machining force) is applied on the
vibrating tool to maintain a certain gap distance between the tool and workpiece (known as
machining gap). The vibrating tool causes the abrasive particles within the machining zone to
impact the workpiece surface causing indentation, microcracks, and eventually material removal.
This process continues leading to the formation of a cavity whose geometry is similar to that of the
tool tip.
Commonly used materials for booster and horn are Monel, titanium alloys (e.g., TiAlV64),
stainless steel, and aluminum alloys (e.g., AlCuMg2) (McGeough 1988). Titanium and aluminum
alloys have high fatigue strengths, and hence they can be used for high vibration amplitudes of up to
40 mm at 20 kHz vibration frequency. The tools used in USM process are usually made of materials
with high wear resistance and fatigue strength properties (Thoe et al. 1998). Tool materials such as

Page 14 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 16 Cause-and-effect diagram for USM

tungsten carbide, silver steel, pure tungsten, and copper are commonly used for machining of the
materials with low fracture toughness such as glass, while chromium-nickel steel is recommended
for machining of the materials with higher toughness such as sintered carbides (McGeough 1988).

Process Parameters and Performance Measures


Ultrasonic machining is a free abrasive machining process where abrasive particles are in interaction
with both the tool and workpiece, thereby causing some levels of intricacy in the process, which is
further influenced by USM process parameters. These parameters are generally controllable input
factors affecting the machining conditions. The machining conditions in turn determine the outcome
of the machining process assessed by performance measures. The process parameters are presented
in Fig. 16 using a cause-and-effect diagram.
Performance measures in USM process include surface integrity, material removal rate (MRR),
and tool wear rate (TWR). Surface integrity, as a crucial performance measure in components made
of hard and brittle materials, comprises surface roughness, subsurface damage, and heat-affected
zone (HAZ). The material removal in USM process involves crack initiation and propagation
followed by chip breakage resulting in a rather coarse surface with shallow pits and voids and
with subsurface microcracks which are generally not acceptable for delicate structural components.
In contrast, HAZ is of lesser concern in USM since the process is considered as a nonthermal process
which generates negligible amount of the heat. Thus, the machined surface is generally free from
thermally damaged zone or residual stress (Thoe et al. 1998).
MRR is considered as a primary process measure which reflects the efficiency of the machining
operation. Besides the volumetric machining rate, drilling speed which is the penetration speed of
the tool into the workpiece may also be used to measure the MRR in USM process. MRR might not
be constant throughout the machining process due to the variation in the machining zone conditions
such as distribution and speed of the free moving particles, indentation force introduced to the

Page 15 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

abrasive particles, intensity of the cavitation bubble collapse, interparticle collisions, and flushing of
the accumulated debris.
Tool wear in the USM process consists of longitudinal wear (WL), side wear (WD), and cavitation
wear (Adithan 1974). The length and weight of the tool decrease as a result of the tool wear, thereby
causing the tool to vibrate out of the resonant frequency which results in a mismatch with the output
frequency of the generator. This leads to reducing amplitude and subsequently a decreased MR-
R. Furthermore, tool profile may change as a result of tool wear, thereby affecting the dimensional
accuracy of the machined features. TWR in USM is influenced by parameters such as particles’ type
and size, workpiece and tool material, tool size, and rate and method of slurry delivery to the
machining gap (Adithan 1974).

Capabilities and Applications of USM


USM processes are capable of machining a wide range of workpiece materials regardless of their
electrical or chemical characteristics. Materials can be categorized into three groups in view of their
machinability in USM process. Type I materials such as glasses are very brittle and hence machined
easily by the USM process. The material is removed as a result of the propagation and integration of
the minute cracks originated from the interactions of abrasive particles with surface material. Type II
materials, like hardened steels, exhibit some plastic deformation before fracture, and they can be
machined by USM process but with a lower efficiency and MRR as compared to Type I materials.
Type III materials such as copper and soft steel are ductile with a high fracture toughness, and hence
they are unsuitable for machining by USM technique. The applications of the USM process include:

• Various machining operations such as drilling, die sinking, and contour machining of features
with circular and noncircular geometries as well as machining of the complex shapes and 3-D
contours
• Machining of ceramics, glasses, silicons, germaniums, quartzes, sapphires, ferrites, optical fibers,
and sintered carbides (Type I) as well as hardened stainless steels, hard carbon alloys, and nickel-
titanium alloys (Type II)
• Application of USM in micromachining (i.e., termed micro-USM) for microhole drilling, slot
machining, and -D microcavities milling

The advantages of the USM process are in the following:

• USM is an efficient and cost-effective technique for precision machining of hard, brittle, and
fragile materials especially in medium or small quantities.
• USM is an environment-friendly process, suitable for machining both conductive and
nonconductive materials.
• Since actual machining is carried out by abrasive particles, tool materials used can be softer than
that of the workpiece material.
• Unlike other thermal-induced processes, USM does not create heat-affected zones, which may
induce residual stresses in the machined surface.

Micro-Ultrasonic Machining (Micro-USM)


Micro-USM technique was introduced by Masusawa’s group in the mid-1990s (Sun et al. 1996a).
Since then, a number of research works have been carried out to enhance this process in both
technological and fundamental aspects. The capabilities of micro-USM technique have been

Page 16 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 17 Machining of microfeatures using die-sinking and contouring mode micro-USM (Medis and Henderson 2005;
Boy et al. 2010) (a) An array of pillars in PZT with diameter of 280 mm and depth of 6,000 mm. (b) Micro channels with
serpentine pattern on silicon

demonstrated for various micromachining operations such as drilling of microholes, machining of


straight and spiral grooves, machining of complex-shaped (3-D) microstructures, and generating of
microfeatures in MEMS components (Medis and Henderson 2005). These machining operations
have been carried out by implementing two modes of micro-USM, namely, die-sinking mode and
contouring (scanning) mode.
In die-sinking mode, micro-tool is fabricated with a face containing the patterns similar to features
required in workpiece. During the machining, micro-tool is driven toward the abrasive slurry and
workpiece, whereby the patterns can be transferred from the tool onto the workpiece. This mode of
micro-USM is capable of generating either a single complex microstructure or a pattern of simple
microfeatures in one pass of the tool penetration toward the workpiece. In contouring mode,
typically a micro-tool with simple shape such as a cylindrical rod is used. The desired microfeature
can be machined by guiding the micro-tool along a predetermined tool path. Figure 17 depicts the
microfeatures machined on different materials using die-sinking and contouring mode of micro-
USM.
Since the inception of micro-USM in the mid 1996s, there have been some developments in the
machine system design and configuration including tool preparation and tooling system, tool-
workpiece interface monitoring, force measurement and monitoring, workpiece holding and horn
design. These advancements have resulted in improvements to process efficiency and machined
surface quality in micro-USM.

Comparison of USM and Micro-USM


The process conditions under which material removal takes place in USM and micro-USM are
distinctively different due to the diverged process parameter magnitudes employed for
micromachining. In order to satisfy the micromachining requirements, the mechanical stress causing
crack initiation, propagation, and eventually fracture of the workpiece material is applied to a very
small area or volume of the workpiece to reduce the unit removal and to attain low surface roughness
(Masuzawa 2000). This is typically achieved by downscaling the particle size with magnitudes of
a few hundreds of micrometers used in USM to micrometer or even submicrometer range suitable
for micromachining (Egashira and Masuzawa 1999). Another factor that contributes to reduction of
the unit removal for micro-USM applications is introducing smaller energies into the slurry medium

Page 17 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Table 1 Process parameters and machining conditions in USM and micro-USM (Thoe et al. 1998; Zhang et al. 2005;
Hu 2007)
Process parameters and conditions USM Micro-USM
Vibration frequency Usually  20 kHz Usually  20 kHz
Vibration amplitude Tens of micrometers (5–50 mm) Within micrometers (0.2–5 mm)
Abrasive particle Size Tens of micrometers (50–300 mm) Within micrometers (0.2–5 mm)
Machining load 0.1–30 N Within 100 mN
Tool size Usually  1 mm Within 500 mm
Vibrated component Tool Tool or workpiece
Slurry delivery and flushing method Jet flow, suction, or combined method Slurry bath and tool rotation

and consequently to each single particle in the machining zone as compared to that of USM. This
leads to a reduced indentation depth and eventually a smaller removal volume per particle impinge-
ment in the process. This requirement is achieved by lowering the vibration amplitude to the range of
a few micrometers (Zhang et al. 2005). Machining conditions and magnitudes of the process
parameters in USM and micro-USM are presented in Table 1.
Further to above conditions desired to satisfy micromachining requirements, and from the
standpoint of the feature size magnitudes in micromachining, a tool with dimensions in the
micrometer range is required in micro-USM process. This requirement confines using methods
such as jet flow and/or suction flushing which are normally employed in USM, and instead it favours
the use of tool rotation to facilitate the slurry circulation and debris removal in the machining zone
(Egashira and Masuzawa 1999). The flushing conditions in USM are affected by flow rate and flow
pressure while the flushing conditions in micro-USM are controlled by the centrifugal forces
resulted from the tool rotational speed. Therefore, these factors give rise to variations in conditions
of the tool-workpiece interface between USM and micro-USM.
Nevertheless, since micro-USM is an adoption of the conventional USM, both processes still
share the same principle for machining. Similar to USM, the material removal in micro-USM is
performed by mechanical action of the abrasive particles driven directly or indirectly by either
vibrated micro-tool or workpiece as well as by cavitation in the slurry fluid causing the particle
indentation, microcrack initiation, propagation, and fracture breakage in the impacted material
(Egashira and Masuzawa 1999). This implies that certain process performance results are common
for the two processes. Material removal is implemented by multiple-particle indentations and
impacts, which cause the machined surface to be covered by numerous minute craters with randomly
distributed positions. This effectively limits the minimum surface roughness achievable in the
process. Also, the machined workpiece surface may contain a layer with many microcracks which
is not acceptable for the majority of microstructures and microcomponents (Brinksmeier et al. 1998).
Although these surface quality issues are encountered in both USM and micro-USM, their influence
is more significant in micro-USM due to relatively smaller size difference between surface defects
and machined microfeatures. Moreover, the surface roughness should be reduced in proportion to
the feature size as an important factor for micromachining. The ability to control and predict the
micro-USM process performance is also more crucial due to miniature tools, small particles, and
features size involved in the process. In addition, there is a need to improve the accuracy and
reliability of the micro-USM system, which is considered as an important factor to perform
micromachining. Thus, improvements to the micro-USM technique require advancements that
provide better understanding of material removal mechanisms and conditions, better control and

Page 18 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

prediction of process performance, as well as better precision and reliability of the machining
system.

Configuration of the Machine System


Two major configurations have been proposed for micro-USM system, namely, tool vibration and
workpiece vibration methods. In the former method, the tool is attached to ultrasonic horn and it is
oscillated in a vertical direction, while workpiece is fixed onto the XY stage (Sun et al. 1996b). Also,
a rotational movement is applied to the horn and attached tool through a spindle system. While
workpiece clamping is simple in this type of micro-USM system, implementing the vibration unit
into the tooling system to some extent limits the rotational accuracy of the spindle system which is
undesirable from the standpoint of the requirements for a high-precision micromachining process
(Masuzawa 2000).
In the workpiece vibration method, the oscillation of the tool is removed; thus tooling system is
only responsible for guiding and rotating the micro-tool. Applying the vibration to workpiece rather
than tool is an attractive method due to its simplicity in the design and operation of the spindle
system, precision of the tooling system, effective agitation of the slurry, and efficient delivery of the
particles into the machining zone (Egashira and Masuzawa 1999). Furthermore, unlike the micro-
USM with tool vibration method, tool wear has no influence on the vibration amplitude of tool tip in
workpiece vibration method. Despite these advantages, issues such as reliable workpiece clamping
and consistent transmission of the vibrations from horn to the workpiece need to be addressed to
improve the process performance.

Tool Preparation and Tooling System


Miniaturization of the tool imposes some constraints on implementation of the micro-USM process,
particularly in the fabrication of tiny tools and configuration of tooling system (Zhang et al. 2005).
Mounting and aligning of the micro-tools are key issues, which have a direct impact on process
productivity and part accuracy. Various tooling strategies and tool preparation techniques such as
on-machine tool preparation using wire electrical discharge grinding (WEDG) and ultrasonic-
assisted grinding, utilizing the sintered diamond tools, and fabrication of the multi-tools have
been proposed to enhance the capability of the micro-USM process. Also, rotation of the micro-
tool has been reported to reduce the form error and to improve the machining rate (Egashira and
Masuzawa 1999). However, tool rotation has the adverse effect of enlargement of the microhole
diameter mainly due to the spindle runout. While improvements are made to rotational and linear
movements of the micro-tool, the flexibility of the tooling system also should be considered to
accommodate the tool preparation techniques.

Monitoring and Control of the Machining Force


During material removal in micro-USM process, the micro-tool progresses further toward the
workpiece through abrasive slurry to maintain a certain machining gap and thus a stable machining
condition. While a machining gap larger than an optimum level would cause the particles to lose
their effective kinetic energy before striking the workpiece surface, using a very small machining
gap might result in suppressing the abrasive particles by the micro-tool and hence leaving them with
insufficient energy for impacting the workpiece surface. In both conditions, the material removal
rate would be very low or even negligible. Furthermore, after the onset of the machining process, the
interface boundary between tool and workpiece as well as the conditions of machining zone vary
continuously, and hence it is difficult to maintain the machining quality (Zhang et al. 2005). The

Page 19 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 18 Schematic of the developed micro-USM system

control of the machining gap in micro-USM is made possible through precision measurement and
control of the static load (machining force) as the feedback in the process.
While in USM, static load and tool size are in the range of a few kilogram-force and larger than
1 mm, respectively; in micro-USM process the static load is in order of a few gram-force and tool
size ranges within 500 mm (Zhang et al. 2005). Besides, mismatched tool-workpiece interactions not
only adversely affect the stability of the micro-USM process, but they also lead to the breakage of
micro-tools. About 60 % of the micro-tool breakages have been reported to occur due to mishandling
of the initial engagement of tool-workpiece interaction (Zhang et al. 2006). Therefore, downscaling
of the process parameters from USM to micro-USM necessitates the development of a reliable force
sensing and monitoring system with high precision and rapid response in order to enhance the
stability of the micromachining process and to avoid tool breakages.
Different strategies have been proposed to control the machining gap and contact force between
micro-tool and abrasive slurry in micro-USM process. Acoustic emission (AE) technique was
proposed and implemented to monitor the tool-workpiece interaction in order to reduce the
probability of the tool breakage in the process (Moronuki and Brinksmeier 2002). However, the
AE output signal has been reported to be influenced by the workpiece position on the machine table
as well as the intrinsic characteristics of the ultrasonic generator used in the process, thereby causing
an incorrect assessment on actual state of the tool-workpiece interaction and hence incomplete
machining (Zhang et al. 2006).

Page 20 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 19 Ultrasonic horn booster with integrated vacuum chuck for workpiece holding

Other methods such as using electronic balance and dynamometer also have been proposed for
measuring the machining force in micro-USM. A digital balance with resolution of 10 mg and
response time of 10 ms was proposed in (Sun et al. 1996a). Also, an electronic balance with
a minimum index of 1 mg was installed in a micro-USM experimental setup acting as the sensor
for feedback control (Egashira and Masuzawa 1999). However, low-frequency response and small
sampling rate were identified as major limitations of this method. In another study, a precision force
actuation and measurement system with the capability of sensing and control of forces in the low mN
range was developed (Hoover and Kremer 2007). Nevertheless, the force measurement in this
system only provides the desired accuracy when employed for measurements in the horizontal
direction rather than vertical one. Thus, it may not be suitable for ultrasonic machining process.

Workpiece Holding Method


The machining accuracy is directly affected by the workpiece holdingmechanism in micro-USM
with workpiece vibration method. In such a setup, the workpiece must be attached firmly to the
ultrasonic horn, which otherwise mechanical vibrations would not be transmitted to the workpiece
consistently, causing unstable machining conditions and excessive heat generation in the horn-
workpiece interface.
The use of double-sided adhesives to attach the workpiece to the transducer or horn has been
attempted by researchers (Hu 2007). However, this method has two main drawbacks. The removal
of workpiece after machining is difficult, especially in the case of thin and fragile workpiece
materials. Also, the adhesive tape itself may absorb the ultrasonic power and alter the acoustic
characteristics of the original mechanical vibration supplied into the ultrasonic horn. Therefore,
a reliable clamping system is deemed necessary to attain a uniform transmission of the mechanical
vibrations from ultrasonic horn to the workpiece.

Development of a Micro-USM System with New Techniques for Workpiece Clamping and
Static Force Control
A new micro-USM system was designed and fabricated with the purpose of improving the
functionality and flexibility of the machine system (Zarepour et al. 2011). Effective and well-
proved techniques such as applying the vibration to workpiece are incorporated in developed
system. Figure 18 illustrates the schematic diagram of the developed micro-USM system which
consists of five main subsystems, namely, generation and transmission of ultrasonic vibrations,

Page 21 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 20 Configuration of the force measurement system

tooling system, workpiece holding, measurement and control of static load, and slurry delivery
system.
The tool rotation has been adopted to reduce form errors and to provide flexibility in the tooling
system. The whole tooling system is mounted on a three-axis stage. The micro-tool motion is
controlled with a minimum incremental motion of 0.03 mm in z-axis and 10 mm in both x-axis and
y-axis. An integrated piezo-motor controller/driver with computer interface is used in z-axis,
enabling a highly reliable tool positioning with 30 nm sensitivity. This configuration provides an
accurate control over the infeed of the tool in the direction of the workpiece oscillation in order to
maintain a preset machining gap.

Workpiece Clamping System The task of holding the workpiece against the ultrasonic horn
hampers the efficient use of the micro-USM with workpiece vibration method. The vibration
frequency of the horn is 50 kHz in the developed system. As such, attachment of the workpiece to
the horn is a critical stage in the process, which may affect the stability and accuracy of the
machining process. The micro-tool used has a diameter up to 300 mm, which is susceptible to
breakage in the event of tool collision.
An aluminum booster together with a full-wave titanium horn, which has a recess to accommo-
date the workpiece, is used to transmit the vibration from ultrasonic transducer to the workpiece.
Nominal vibration amplitude of the system ranges between 0.8 and 5.5 mm using a combination of
the reverse booster and horn with the same gain ratio of 0.5:1 and adjustment of the output power of
the ultrasonic generator. As shown Fig. 19, a vacuum chuck is incorporated to the ultrasonic horn for
workpiece clamping. It is important to introduce the vacuum tube to the ultrasonic horn at the
position of the nodal point where the amplitude of vibration is close to zero. The vacuum tube is
connected to the vacuum pump through a liquid separator with filter. The proposed method enables
a rapid clamping and unclamping of the workpiece without introducing crack to thin and fragile
workpieces.

Measurement and Control of the Static Load In the developed system, the static load between
micro-tool and vibrated workpiece is measured and controlled using a precision force sensor
integrated with the tooling system as illustrated in Fig. 20. The proposed design has a merit in
that the sensor can be mounted on tooling system without hampering the rotation of the micro-tool.

Page 22 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Table 2 Mechanisms of the material removal in ultrasonic machining


Mechanism of
material removal Description References
Direct hammering Material removal occurs through direct localized hammering of abrasive Thoe et al. (1998)
particles by tool against the workpiece surface. Only the larger particles
take part in the mechanism either because they rest on the workpiece,
thus receiving a blow from the tool face, or because they descend in
contact with the tool face to hammer the work. This mechanism may
result in crushing of the abrasive particles
Free impact Material removal is performed by the impact of free moving abrasive Soundararajan and
particles on the workpiece. This mechanism generally takes place when Radhakrishnan (1986)
the particles with a size smaller than the gap distance fly across the
machining zone and strike the workpiece surface at random locations
Cavitation erosion Erosion of the surface occurs due to cavitation effects within the Soundararajan and
abrasive slurry. This phenomenon is considered as a form of wear Radhakrishnan (1986)
whereby the damage is produced by bubbles collapsing near the
surfaces of both workpiece and tool
Chemical action This mechanism is associated with the fluid employed for the slurry and Thoe et al. (1998)
involves some chemical reaction taking place within slurry mainly due
to micro-explosion resulted from cavitation in the slurry fluid

Furthermore, utilizing the sensor in the tool side eliminates the measurement errors arising from
dead weight of the ultrasonic stack and fixture as well as the noise and vibration from ultrasonic
horn. The sampling frequency of the sensor is up to 1 kHz which can be utilized in both tension and
compression; hence, it is capable of high-speed measurement of the variation in the machining force.
The static load can be constantly monitored by the force sensor in order to maintain the preset
machining gap through controlling the infeed tool motion. The system also has a built-in overload
protection feature.

Material Removal Mechanisms and Modes in USM and Micro-USM


The mechanisms of material removal in USM process can be categorized into four types, namely,
direct hammering, free impact, cavitation erosion, and chemical action. Detailed descriptions of
these mechanisms are presented in Table 2. One material removal mechanism or a combination of
mechanisms will dominate the material removal process depending on the process parameters (Thoe
et al. 1998). While direct hammering and free impact are primarily responsible for material removal
in USM, cavitation erosion and chemical action are of secondary significance with the majority of
the workpiece materials such as glass, ceramics, and hardened steel (Thoe et al. 1998). Cavitation
erosion has been reported to play an important part in machining of the porous materials (Weller
1984).
The effect of these mechanisms results in material removal from the workpiece either by crack
formation and fracture in brittle mode or by cutting and shearing in ductile mode. In addition, during
the USM process, there might be a situation where the material at transient surface of the workpiece
is not completely removed but only displaced upon impingement of the abrasive particles.
When material is impacted by a hard angular particle at high speeds or under large contact forces,
plastic deformation occurs in the contact zone due to high compressive and shear stresses, and
a radial crack is formed. After impact, the plastic deformation results in large tensile stresses leading
to initiation of lateral cracks and eventually material removal (Wensink and Elwenspoek 2002).
When the speed of particles or the amount of force acting on them during the process is below the

Page 23 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

threshold values required for crack initiation in the brittle material, particles may cause only plastic
deformation to the surface, which brings about material removal by ploughing and cutting in
a ductile mode instead of a brittle one. This change in erosion mode from brittle fracture-dominated
behavior to plastically dominated behavior is called brittle-ductile transition (Wensink and
Elwenspoek 2002). The material removal mode, whether it occurs in brittle or ductile mode, and
transition between these modes have a direct effect on both surface integrity and material removal
rate associated with process quality and productivity, respectively.
Although the underlying machining principle of micro-USM is similar to that of conventional
USM (Zhang et al. 2005), downscaling of the process parameters such as tool size, vibration
amplitude, machining load, and particles size so as to minimize the contact zone of abrasive particles
with tool and workpiece to the microscale range may affect the machining condition. Also, the
nature of material removal in micro-USM, whether the material is removed through ductile
deformation or crack generation followed by brittle fracture, may differ from that in USM due to
the influence of miniaturized tool-abrasive-workpiece interaction intensity (Zhang et al. 2006). For
instance, the accumulation of workpiece debris and crushed abrasives in a small machining gap and
difficulty in removing them from the machining zone due to susceptibility of the micro-tool to
deformation and breakage will affect the manner by which material is removed from the workpiece
surface (Yu et al. 2006). This also could lead to further complexity in tool-abrasive-workpiece
interactions and consequently more intricate material removal mechanisms in micro-USM.
Research works concerning the material removal modes in micro-USM are very limited in the
literature, while no studies have been reported regarding the investigation on dominant material
removal mechanisms that contribute to the machining process.

Intricacies Involved in Study of Material Removal Mechanisms in USM and Micro-USM


Characteristics of the machined surface are dependent on the manner by which the material is
removed from the workpiece. The predominant mechanism of material removal has a direct
influence on material removal rate and machined surface quality. The study of different conditions
in tool-abrasive-workpiece interaction within the slurry fluid is complex due to the following issues:

• Ultrasonic machining is a free abrasive machining process. That is, abrasive particles move
randomly and freely within the machining zone during the material removal process. Thus, the
location of each particle is changing by instant displacement and rotation within the slurry.
• Generally, the shape of the particles is irregular and with a distributed size over the machining
zone. That is, the real particles engaged in machining operation do not have exactly the same size,
but they follow a nominal average size. Besides, abrasive particles are not ideally rigid and
incompressible; thus, they break into smaller pieces due to collision with the workpiece and tool
or between themselves resulting in variation of the abrasive size and distribution across the
machining gap.
• The distribution of the abrasive particles in the affinity of the machined surface and tool may
become uneven as a result of slurry flow and agitation due to micro-explosion of the bubbles
inside the slurry fluid. Also, there is a likelihood that the conditions of the machining gap become
abnormal due to embedding of the particles into the workpiece or tool surface (Yu et al. 2006).
These factors may lead to a situation whereby the mechanical vibration would not be transmitted
to the slurry fluid and abrasive particles effectively.

An eclectic of these factors introduces micro-USM as a stochastic process. Further, as discussed


earlier in this chapter, the effective material removal mechanisms may be different in the case of

Page 24 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 21 Material removal mechanisms involved in micro-USM process: (a) pure cavitation; (b) direct hammering; (c)
free particle impact

micro-USM as compared to that of conventional USM due to more complexity of the tool-abrasive-
workpiece interaction in the machining zone.

Investigation on Material Removal Mechanisms in Micro-USM


Experimental observations by authors (the results to be published in a separate paper) demonstrated
that the material removal in micro-USM is primarily caused by the erosive wear associated with
ultrasonic vibration rather than by abrasive wear resulting from micro-tool rotation during the
process. The erosive wear in micro-USM mainly takes place via three material removal mechanisms
similar to those of conventional USM as described earlier in this chapter. These mechanisms include
pure cavitation due to bubble implosion inside the slurry liquid, direct hammering of abrasive
particles on the workpiece surface by micro-tool, and impact of the free moving abrasive particles.
While the role of pure cavitation has been reported to be insignificant in USM, direct hammering and
free particle impact are regarded as main removal mechanisms. Also, free particle impact has been
confirmed to account for only about 22 % of total material removal in USM (Khairy 1990). Thus,
direct hammering has been identified as the primary cause of the material removal in conventional
USM. The relative contribution of the removal mechanisms might be different in micro-USM as
compared to that of conventional USM due to the diverse removal conditions inside the machining
gap for two processes. As such, a basic study is conducted to identify the contribution of the different
mechanisms to material removal in micro-USM. These mechanisms are depicted in Fig. 21.

Material Removal by Pure Cavitation To identify the role of pure cavitation in material removal,
a set of machining experiments were conducted by using deionized water and without any abrasive
particles. Tungsten rods with a diameter of 300 mm and monocrystalline silicon with thickness of
525 mm were used as micro-tool and workpiece, respectively. The vibration frequency of 50 kHz and
vibration amplitude of 2 mm were used for all experiments. A machining duration of 5 min was
maintained for all experimental runs, and each run was repeated three times. The force introduced by
ultrasonic vibration was recorded using the precision force sensor with interface software. This force
value corresponds to the adjusted gap distance, and it is used as the feedback to maintain the gap
distance during the process. Figure 22 shows the profile of a typical microhole with maximum depth
of 6.8 mm obtained at gap distance of 2 mm.

Material Removal by Particle Direct Hammering Another set of experiments was planned and
conducted to investigate the role of direct hammering mechanism in material removal in the micro-

Page 25 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 22 Material removal resulted from cavitation only (without applying abrasive particles); vibration frequency ¼ 50
kHz, vibration amplitude ¼ 2 mm, gap distance ¼ 2 mm, machining time ¼ 5 min

Fig. 23 Material removal resulted from direct hammering of the particles by micro-tool; vibration frequency ¼ 50 kHz,
vibration amplitude ¼ 2 mm, gap distance ¼ 0.2 mm, abrasive particles ¼ PCD, particles size ¼ 3 mm, machining
time ¼ 5 min

USM process. PCD particles with nominal size of 3 mm mixed with DI water were utilized as
abrasive slurry. Vibration amplitude was set at 2 mm, and small machining gaps in the range of
0.2–2 mm were applied. This is to ensure that abrasive particles can be pushed and impacted directly
by end face of the micro-tool onto the workpiece surface, that is, the condition required for direct

Page 26 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 24 Material removal resulted from free particle impact; vibration frequency ¼ 50 kHz, vibration amplitude ¼ 2.0
mm, gap distance ¼ 5 mm to 200 mm, abrasive particles ¼ PCD, particles size ¼ 3 mm, machining time ¼ 5 min

hammering mechanism to take place in micro-USM process. Other machining conditions were
similar to that described for the experiments in the previous experiment.
The depth of microholes was measured, and maximum machined depth of 28.4 mm was obtained
when machining gap was adjusted at 0.2 mm. The profile of a typical microhole machined under
these process conditions is depicted in Fig. 23.

Material Removal by Free Particle Impact Machining experiments were conducted to investi-
gate the effect of free particle impact in the material removal process. PCD particles with nominal
size of 3 mm mixed with DI water were utilized as abrasive slurry. Vibration amplitude was set at
2 mm, and machining gap was adjusted in the range of 5–200 mm. Since the machining gap was set to
be larger than the particles size, direct hammering of the particles by tool face was unlikely to occur
in the process. Therefore, material removal was brought about primarily by the impact of the free
moving particles striking the workpiece surface.
The depth of microholes machined at various gap distances was measured and analyzed. As
shown in Fig. 24, a microhole with a depth of approximately 218.2 mm was obtained, which
corresponds to the microhole machined at 30 mm gap distance.

Contribution of Different Material Removal Mechanisms in Micro-USM A comparison of the


three removal mechanisms in micro-USM with respect to their contributions to material removal is
presented in Fig. 25. Average machining depth was calculated for three machining runs at which the
maximum depth was achieved under different mechanisms in the process. While pure cavitation
produced a shallow microhole with an average depth of only 6.8 mm during 5 min of the machining
process, direct hammering and free particle impact produced microholes with an average depth of
28.4 mm and 218.2 mm, respectively, over the same machining time. Accordingly, the depth of the
microholes produced under the free particle impact mechanism is approximately 32 times and
8 times higher than that of pure cavitation and direct hammering, respectively. This could be
attributed to high velocity of the free moving particles inside the machining gap which are subjected

Page 27 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Fig. 25 Comparison of different removal mechanisms with regard to their role in micro-USM

Fig. 26 Approach to developing the predictive model for material removal mode in micro-USM

to ultrasound in micro-USM. Results of this study evidence that unlike conventional USM, free
particle impact is the main contributor to material removal in micro-USM.

Predictive Modeling of Ductile and Brittle Removal Modes in Micro-USM


The results of the investigation on abrasive particle erosion of some brittle materials have demon-
strated that the ductile-brittle transition may occur as the size of particles is reduced (Hutchings
1992). Since the size of abrasive particles used in micro-USM is smaller than that of conventional
USM by a factor of 10 or more, the likelihood of such a ductile-brittle transition is higher in micro-
USM as compared to that of conventional USM. Being a loose abrasive process, the material
removal mode in micro-USM can be influenced by setting the process parameters including particle
size at desired levels in order to enhance the process performance.
Investigations on material removal characteristics have been carried out based on the morphology
of surfaces, processed by multiple-particle impingements. Unlike the multiple-particle impact, the
method of single-particle impact has a potential to provide more basic and explicit information about
material removal modes and mechanisms. Investigation on craters produced by single-particle
impact is an attractive approach with the capability of providing practical solutions to overcome

Page 28 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

the ambiguity around the issue of determining the material removal modes in abrasive-based
micromachining processes. Therefore, the study of single abrasive particle impingement may
open new avenues toward a more fundamental and objective approach to investigate the brittle
and ductile machining modes and transition between them in micro-USM process.

Approach to Development of the Predictive Model


Figure 26 outlines a model for prediction of material removal mode in micro-USM process. The
model consists of three parts: estimation of the apparent threshold kinetic energy for radial and
lateral fracture in workpiece material, estimation of the kinetic energy of an impinging particle inside
the machining zone, and criteria for determining the material removal mode. The analysis to estimate
the threshold kinetic energy is performed based on the well-established indentation fracture theory
for microcracks initiation by hard angular particles in work materials. The application of this theory
to predict the thresholds for ductile and brittle transitions in other processes such as particle erosion
have been described by Hutchings (Hutchings 1992).
The model estimates the kinetic energy of a single impinging abrasive particle driven by the
ultrasound in the process. The primacy on the type of material removal mode lies at the foundation of
comparative principle between the apparent threshold kinetic energy of the workpiece material and
estimated kinetic energy of the impacting particle based on the input parameters of micro-USM
process.

Summary
The major discussions and findings presented are summarized in the following:

• The development of a micro-USM system has been reported, which incorporates the workpiece
vibration method. New techniques for the force measurement and workpiece clamping in micro-
USM were introduced.
• A study was performed on the role of various mechanisms contributing to the material removal in
the process. Results demonstrated that the depth of microholes machined by free particle impact
mechanism is larger than that of microholes machined under pure cavitation and direct hammer-
ing mechanisms by 32 times and 8 times, respectively. Therefore, the free particle impact was
found to play a major role in the material removal in micro-USM process.
• A new model was developed to predict material removal mode in micro-USM with respect to
material properties of the workpiece and particles as well as ultrasonic characteristics of the
system. In the model development, the required conditions for ductile and brittle removal in
micro-USM were discussed. The kinetic energy of a single impinging abrasive particle introduced
by the ultrasound was estimated in the model. The type of material removal mode was determined
based on the comparison between the apparent threshold kinetic energy of the workpiece material
and estimated kinetic energy of the impacting particle. Experiments were conducted on quartz
material to validate the capability of the proposed model. Modeling results and experimental data
showed that the proposed model is able to predict the material removal modes reasonably well in
the micro-USM process.
• Process models were proposed to predict the material removal rate for brittle and ductile removal
modes in micro-USM. The models were based on the indentation fracture theory and formation of
the plastic indentation zone as well as lateral cracks during the interaction between sharp particles
with workpiece surface. Machining experiments were carried out using the developed micro-
USM system in order to validate the models. MRR values predicted by the proposed models
showed a reasonable agreement with the experimental trends of the MRR for quartz material.

Page 29 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

However, it was observed that the models consistently overestimate the measured MRR values
within the machining conditions carried out in the work. The deviation between theoretical and
experimental results may be due to conditions such as overlapping of the craters, size and shape
distribution of particles, and interparticle collision within the machining zone.

Summary
To date, numerous process principles have been developed to materialize the loose abrasive
machining process in different ways. These developments have been made in response to the
changing needs of the commercial market and to the improvement of the quality of products.
Loose abrasive machining processes are also used for deburring, edge finishing, and surface
integrities. Moreover, process modeling and simulation of the loose abrasive machining are being
studied intently. Although this chapter cannot cover all research and development related to loose
abrasive machining, the authors hope that this chapter will be of assistance in understanding loose
abrasive machining and in supporting future technology innovation.

References
Adithan M (1974) Tool wear studies in ultrasonic drilling. Wear 29:81–93
Balamuth LA (1945) Method of abrading. Patent 602801
Boy JJ, Andrey E, Boulouize A, Khan-Malek C (2010) Developments in microultrasonic machining
(MUSM) at FEMTO-ST. Int J Adv Manuf Technol 47:37–45
Brinksmeier E, Preuss W, Schmutz J (1998) Manufacture of microstructures by ultrasonic lapping.
In: Proceedings of the 13th annual meeting of the american society for precision engineering,
St. Louis
Egashira K, Masuzawa T (1999) Micro ultrasonic machining by the application of workpiece
vibration. CIRP Ann 48:131–134
Gillespie LRK (2006) Mass finishing handbook. Industrial Press, New York
Hoover D, Kremer D (2007) Milli Newton force measurement and control for micro ultrasonic
machining. In: Proceedings of the 15th international symposium on electromachining, Pittsburgh
Hu X (2007) Mechanism, characteristics and modeling of micro ultrasonic machining. PhD thesis,
The University of Nebraska, LN, 182 pp
Hutchings IM (1992) Ductile-brittle transitions and wear maps for the erosion and abrasion of brittle
materials. J Phys D Appl Phys 25:A212–A221
Jha S, Jain VK, Komanduri R (2007) Effect of Extrusion Pressure and Number of Finishing Cycles
on surface roughness in magnetorheological abrasive flow finishing (MRAFF) process. Int J Adv
Manuf Technol 33:725–729
Khairy ABE (1990) Assessment of some dynamic parameters for the ultrasonic machining process.
Wear 137:187–198
Ko S, Baron YM, Chae JW, Polishuk VS (2003) Development of deburring technology for drilling
burrs using magnetic abrasive finishing method. In: Proceedings of international conference on
leading edge manufacturing in 21st century, Japan Society of Mechanical Engineers, pp 367–372
Kurobe T, Imanaka O, Tachibana S (1983) Magnetic field-assisted fine finishing. Bull Jpn Soc
Precision Eng 17(1):49–50

Page 30 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Loveless TR, Williams RE, Rajurkar KP (1994) A study of the effects of abrasive-flow finishing on
various machined surfaces. J Mater Process Technol 47:133–151
Masuzawa T (2000) State of the art of micromachining. CIRP Ann Manuf Technol 49:473–488
Matijevic E, Babu SV (2008) Colloid aspects of chemical–mechanical planarization. J Colloid
Interface Sci 320:219–237
McGeough JA (1988) Advanced methods of machining. Chapman & Hall, New York
Medis PS, Henderson HT (2005) Micromachining using ultrasonic impact grinding. J Micromech
Microeng 15:1556–1559
Mori Y, Yamauchi K, Endo K (1987) Elastic emission machining. Precis Eng 9(3):123–128
Mori Y, Yamauchi K, Endo K (1988) Mechanism of atomic removal in elastic emission machining.
Precis Eng 10(1):24–28
Moronuki N, Brinksmeier E (2002) Micromachining of brittle materials by ultrasonic lapping with
tool wear compensation. In: Proceedings of the third EUSPEN conference, Eindhoven
Nakazawa H (2004) Principle of precision manufacturing (trans: Ryu Takeguchi) Oxford University
Press, New York
Papell SS (1965) Low viscosity magnetic fluid obtained by the colloidal suspension of magnetic
particles. US Patent 3, 215, 572
Patrick WJ, Guthrie WL, Standley CL, Schiable PM (1991) Application of CMP to the fabrication of
VLSI circuit interconnections. J Electrochem Soc 138(6):1778–1783
Preston F (1927) The theory and design of plate glass polishing machines. J Soc Glass Technol
11:214–256
Rabinow J (1948) The magnetic fluid clutch. AIEE Trans 67:1308–1315
Raju HP, Narayanasamy K, Srinivasa YG, Krishnamurthy R (2005) Characteristics of extrude honed
SG iron internal primitives. J Mater Process Technol 166:455–464
Rajurkar KP, Levy G, Malshe A, Sundaram MM, McGeough J et al (2006) Micro and nano
machining by electro-physical and chemical processes. CIRP Ann Manuf Technol 55:643–666
Sato T, Yamaguchi H, Shinmura T, Okazaki T (2007) Study of internal magnetic field assisted
finishing for copper tubes with MRF (magneto-rheological Fluid)-based slurry. Key Eng Mater
329:249–254
Sato T, Wu Y, Lin W, Shimada K (2010) Study of dynamic magnetic field assisted finishing for metal
mold using magnetic compound Fluid (MCF). Key Eng Mater 447–448:258–262
Shimada K, Matsuo Y, Yamamoto K, Zheng Y (2008) A new float-polishing technique with large
clearance utilizing magnetic compound Fluid. Int J Abras Technol 1(3/4):302–315
Soundararajan V, Radhakrishnan V (1986) An experimental investigation on the basic mechanisms
involved in ultrasonic machining. Int J Mach Tool Design Res 26:307–321
Sun XQ, Masuzawa T, Fujino M (1996) Micro ultrasonic machining and self-aligned multilayer
machining assembly technologies for 3D micromachines. In: Ninth annual international work-
shop on micro electro mechanical systems, IEEE proceedings: an investigation of micro struc-
tures, sensors, actuators, machines and systems. San Diego, pp 312–317
Sun XQ, Masuzawa T, Fujino M (1996b) Micro ultrasonic machining and its applications in
MEMS. Sensors Actuators A Phys 57:159–164
Thoe TB, Aspinwall DK, Wise MLH (1998) Review on ultrasonic machining. Int J Mach Tools
Manuf 38:239–255
Tricard M, Dumas PR, Golini D, Mooney T (2003) Prime silicon and silicon-on-insulator (SOI)
wafer polishing with magnetorheological finishing (MRF). In: Proceedings of 2003 ASME
international mechanical engineering congress & exposition (IMECE 03), pp 1–10
Venkatesh VC, Izman S (2008) Precision engineering. McGraw-Hill, New York, p c2008

Page 31 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Venkatesh VC, Inasaki I, Toenshof HK, Nakagawa T, Marinescu ID (1995) Observation of polishing
and ultraprecision machining of semiconductor substrate materials. CIRP Ann 44(2):611–618
Weller EJ (1984) Nontraditional machining processes. Society of Manufacturing Engineers, Publi-
cations/Marketing Division, Dearborn
Wensink H, Elwenspoek MC (2002) A closer look at the ductile-brittle transition in solid particle
erosion. Wear 253:1035–1043
Yamaguchi H, Shinmura T, Ikeda R (2007) Study of internal finishing of austenitic stainless steel
capillary tubes by magnetic abrasive finishing. ASME J Manuf Sci Eng 129(5):885–893
Yin S, Shinmura T (2004) Vertical vibration-assisted magnetic abrasive finishing and deburring for
magnesium alloy. Int J Mach Tools Manuf 44(12–13):1297–1332
Yu Z, Hu X, Rajurkar KP (2006) Influence of debris accumulation on material removal and surface
roughness in micro ultrasonic machining of silicon. CIRP Ann Manuf Eng 55(1):201–204
Zantyea PB, Kumara A, Sikderb AK (2004) Chemical mechanical planarization for microelectron-
ics applications. Mater Sci Eng R45:89–220
Zarepour H, Yeo S, Tan P, Aligiri E (2011) A new approach for force measurement and workpiece
clamping in micro-ultrasonic machining. Int J Adv Manuf Technol 53:517–522
Zhang C, Rentsch R, Brinksmeier E (2005) Advances in micro ultrasonic assisted lapping of
microstructures in hard–brittle materials: a brief review and outlook. Int J Mach Tools Manuf
45:881–890
Zhang C, Brinksmeier E, Rentsch R (2006) Micro-USAL technique for the manufacture of high
quality microstructures in brittle materials. Precis Eng 30:362–372

Page 32 of 33
Handbook of Manufacturing Engineering and Technology
DOI 10.1007/978-1-4471-4976-7_10-1
# Springer-Verlag London 2014

Index Terms:
Abrasive flow machining (AFM) 4
Acoustic emission (AE) technique 20
Barrel finishing 11
Brittle removal modes 28
Chemical mechanical polishing (CMP) 2
Drag finishing 12
Ductile removal modes 28
Elastic emission machining (EEM) 4
Lapping 2
Loose abrasive machining 1
Magnetic compound fluid (MCF) 10
Magnetic field assisted finishing (MAF) 6
Magnetic float polishing 7
Magnetic fluids (MFs) 7
Magneto-rheological abrasive flow finishing (MRAFF) 5
Magneto-rheological finishing (MRF) 8
Magneto-rheological fluid-based slurry 9
Material removal mechanisms and modes 23
Micro ultrasonic machining 17, 19, 21, 23, 28
configurations 19
developement of 21
ductile and brittle removal modes 28
machining of microfeatures 17
material removal mechanisms and modes 23
monitoring and control 19
tools 19
vs. USM 17
workpiece holding method 21
Polishing 2
Preston’s equation 1
Single particle impact 28
Ultrasonic machining (USM) 14–16, 23
advantages 16
basic elements 14
capabilities and applications of 16
material removal mechanisms and modes 23
micro-USM 16
see also Micro ultrasonic machiningprocess parameters and perfomance measures 15
working principle 14
Vibratory finishing 11
Workpiece clamping system 22
Workpiece holding method 21

Page 33 of 33

You might also like