You are on page 1of 32

Examining the Role of Cutting Fluids in

Machining and Efforts to Address


Associated Environmental/Health Concerns
D. P. Adler, W. W-S Hii, D. J. Michalek, and J. W. Sutherland
Department of Mechanical Engineering � Engineering Mechanics
Sustainable Futures Institute
Michigan Technological University
Shortened Title for Running Head
Cutting Fluids and Environmental/Health Concerns
Keywords
Cutting fluids, Machining, Environment, MQL
Abstract
Cutting fluids have seen extensive use and have commonly been viewed as a required
addition to
high productivity and high quality machining operations. Cutting fluid related
costs and health
concerns associated with exposure to cutting fluid mist and a growing desire to
achieve
environmental sustainability in manufacturing have caused industry and academia to
re-examine
the role of these fluids and quantify their benefits. This work summarizes the
traditional
purposes of cutting fluids and reports on recent analytical and experimental
research to critically
examine these functions. To minimize or even eliminate the concerns associated with
cutting
fluid usage, several recent and novel approaches have been proposed and are
examined.
Introduction
In 2002, over two billion gallons of cutting fluids were used by North American
manufacturers.
Traditionally, cutting fluids have been widely used in machining operations in
efforts to increase
cooling and lubricity, and as a result enhance tool life, reduce process
variability, etc. However,
over the last decade, it has become apparent that fluid-related decisions have all
too frequently
been based upon industrial folklore rather than knowledge-based quantitative
evidence. Recently
there has been a change in this situation, in part driven by the fact that costs
associated with fluid
use often constitute between 7% and 17% of total production costs, as compared to
4% for
tooling costs (King et al., 2001). Fluid related expenses include the cost of
installing a fluid
supply system, fluid purchase and system maintenance, and discarded fluid (waste)
treatment.
Fluid-related costs are large because high production manufacturing plants
frequently utilize
several cutting fluid reservoirs each containing thousands of gallons of cutting
fluid, and often an
entire reservoir is flushed to clean the system when quality issues arise
(Filipovic et al., 2000) �
certainly, reducing the amount of fluid employed can produce significant cost and
waste savings.
Two relatively recent strategies focused on reducing fluid use are Minimum Quantity
Lubrication (MQL) and dry machining.
There are four primary categories of cutting fluids that differ in terms of their
thermophysical
properties, common process applications, and method of treatment. Straight oils are
made up
entirely of mineral or vegetable oils, and are used primarily for operations where
lubrication is
required (Kalpakjian and Schmid, 2001). Despite being excellent lubricants, they
exhibit very
poor heat transfer capabilities. Soluble oils are mixtures of oil and water and
have increased
cooling capabilities over straight oils and offer some rust protection. Concerns
with using
soluble oils center upon maintenance issues like evaporative losses and bacterial
growth. Semisynthetics
are similar to soluble oils in performance characteristics, but differ in
composition
because 30% or less of the total volume of the concentrate contains inorganic or
other
compounds that dissolve in water. Semi-synthetics have better maintenance
characteristics than
soluble oils, but do contaminate easily when exposed to other machine fluids and
may pose a
dermatitis risk to workers. Synthetics are chemical fluids that contain inorganic
or other
chemicals dissolved in large amounts of water and offer superior cooling
performance.
Maintenance is also not a major issue with synthetics, however, cases of dermatitis
are more
prevalent in workers and the lubrication functionality is weaker than with semi-
synthetics
(IWRC, 1996).
Extensive use of cutting fluids in machining operations leads to a sizeable waste
stream.
Responsible handling of used/waste fluid is needed to avoid the contamination of
lakes, rivers,
and groundwater. Such handling may include the pre-treatment and treatment of
cutting fluid
wastes, but even the most disciplined manufacturer may still have fluid-related
environmental
concerns associated with chip/workpiece fluid carry-off. It is worth noting that
the cost of fluid
pre-treatment/treatment is sometimes higher than the purchase price of the fluid
itself, and since
the treatment is not always totally effective, disposal may lead to inadvertent
water
contamination.
In addition to the environmental challenges of managing a used cutting fluid waste
stream,
cutting fluids also introduce several health/safety concerns. The National
Institute for
Occupational Safety and Health (NIOSH) estimates that 1.2 million workers involved
in
machining, forming, and other metalworking operations are exposed to metalworking
fluids
annually (NIOSH, 1998). Dermal exposure to these fluids represents a health
concern, as does
the inhalation of airborne fluid particulate. The application of cutting fluids
within a machining
operation often produces an airborne mist, and medical evidence has linked worker
exposure to
cutting fluid mist with respiratory ailments and several types of cancer (Mackerer,
1989; Thorne
et al., 1996). This makes the use of cutting fluids a health issue with the
potential of both longand
short-term consequences.
Over the past decade, cutting fluids have been studied extensively to characterize
their relative
benefits and shortcomings in terms of their use within machining processes.
Traditionally,
manufacturers have employed cutting fluids to serve the following functions:
cooling, lubrication,
corrosion inhibition, and chip flushing, and as a result, achieve such benefits as
increased tool
life, improved workpiece quality, enhanced machine tool life, and effective chip
management.
However, many issues surround the use of cutting fluids, including fluid system
maintenance,
fluid pre-treatment/treatment/disposal (Skerlos et al., 2000; Skerlos et al.,
2001), and worker
health/safety concerns. This paper summarizes recent work performed to establish an
improved
understanding of cutting fluid function and to describe the salient mechanisms
associated with
fluid mist formation/behavior. It also presents promising approaches that offer
opportunities for
addressing cutting fluid related concerns while not compromising on process
performance.
Traditional Role of Cutting Fluids
Cutting fluids have traditionally been used in machining operations to lubricate
the chip-tool and
tool-workpiece interfaces, remove heat from the workpiece and cutting zone, flush
away chips
from the cutting area, and inhibit corrosion (Shaw, 1942). While each of these four
functions
can be employed as justification for cutting fluid usage, it is widely believed
that the primary
functions of a cutting fluid are lubrication and cooling. Seminal contributions to
the technical
literature in support of this belief are provided below.
Lubrication
Any study on the lubricating effects of a cutting fluid builds upon an
understanding of the
mechanics and forces involved in a machining process. An early method proposed to
analyze a
metal cutting process was the orthogonal cutting model of Merchant (1941). This
model is based
upon the assumptions that the cutting edge is perfectly sharp, deformation is plane
strain, and
that the stresses on the shear plane are distributed evenly. Their model
characterizes the
deformation geometry via the shear angle,f, which describes the plane on which
shear
deformation occurs. The forces acting on the chip at the rake face of the tool are
balanced by the
force acting on the chip at the shear plane. This allows for the development of a
system of force
equations that can be used to determine characteristics of the process.
Based upon the work of Merchant, Lee and Shaffer (1951) used plasticity theory,
specifically
slip-line field theory, to develop a more sophisticated model to apply to the
machining problem.
Oxley and Hastings (1977) added strain hardening into the slip-line theory and
successfully
applied it to predict cutting forces. The predictive abilities of this model were
shown to be
extremely sensitive to the workpiece material. A major conclusion of slip-line
field modeling is
that specification of rake angle and friction factor do not distinctively determine
the shape of the
chip. This is because more than one field can be constructed, each with a different
chip
thickness and contact length with the tool.
Further studies sought to account for the complicating issues of material behavior,
nonlinear
contact, high temperature, high strain rate, and large strain in metal cutting
modeling/simulation.
A great many efforts have been made to use finite element methods to characterize
the metal
cutting process (Iwata et al., 1984; Strenkowski and Carroll, 1985; Lin and Pan,
1993; Marusich
and Ortiz, 1995). Over the last several decades, much work has concentrated on the
development of mechanistic models to predict cutting forces based upon the method
proposed by
Sabberwal (1960). A reasonable amount of success has been achieved by simulating
some
machining operations, but the method is process dependent, and material state
quantities like
stress, strain, and cutting zone temperature are difficult to obtain.
One of the principal challenges associated with the modeling of machining
operations is the
complexity associated with the work-tool-chip interaction. The tool chip interface
is
characterized by sliding contact between the tool and the workpiece at high normal
pressure and
temperature. The energy that is consumed due to friction is mostly converted into
heat on the
rake face, causing tool temperatures to be high. In order to counteract this
extreme frictional
force, cutting fluids have been used as lubricants in some machining operations.
Shaw et al. (1951) experimentally observed that the cutting fluid does not
lubricate at high
speeds. The possible explanations for this behavior included: chips are carrying
cutting fluid
away too quickly for it to reach the cutting zone and serve as a fluid-film
lubricant and the time
is too short for the fluid to chemically react with metal surfaces to form a solid-
film lubricant.
Cassin and Boothroyd (1965) also found that no lubrication was evident at high
cutting speeds.
They suggested that lubrication occurs at low speeds by diffusion through the
workpiece or that
the extreme pressure additives within the fluid react to form a boundary layer of
solid-film
lubricant.
As noted above, while the primary functions of a cutting fluid are considered to be
lubricating
and cooling, lubrication is the dominant function for only those machining
operations that
employ low cutting speeds, e.g., drilling and tapping. It is to be expected that a
fluid in these
operations would reduce the friction between the chip and the rake face. However,
in drilling
and tapping, a significant amount of friction between chip and tool occurs in
locations other than
the rake and flank faces. Another source of friction results when the chips attempt
to evacuate
through the flutes. The chips rub against the tool and hole wall, and in some cases
the chips clog
the flutes, increasing torque and axial force, increasing tool temperature, and
occasionally
marring the hole wall surface. In these cases, the presence of a cutting fluid can
reduce the
friction between the chips and tool flutes, enabling the smooth evacuation of chips
from the hole
and avoiding chip clogging. Of course, the efficacy of the fluid as a lubricant is
very dependent
on the success achieved in delivering it to the bottom of the hole. Furthermore,
the character of
the chips produced in drilling and tapping play an important role in the chip
clogging
phenomenon (Haan et al., 1997; Cao and Sutherland, 2002).
To characterize the clogging phenomenon in drilling requires an understanding of
the chip
formation process. Kahng and Koegler (1976) gave an explanation of chip forming in
twist
drilling. They proposed that chip curl is formed in the shear zone, influenced by
the contact
between the chip and rake face of the tool. Chip breakage resulted when the chip
impacts the
workpiece or the tool. Chip breaking methods were suggested to allow for easier
chip
evacuation. Nakayama and Ogawa (1978) studied the basic formation of chips,
specifically
those formed in the drilling process. Continuous chips were found to have the rake
face of the
tool as a tangential plane to the surface of the chip. Helical chips were described
by their chip
flow angle, radius of upcurl, and radius of sidecurl. The radii are functions of
the cutting
velocity and the angular velocity of the chip. Haan et al. (1997) found that the
significant
variables in determining chip size were the feed, workpiece material, and drill
type.
In summary, for low cutting speeds such as found in operations like drilling and
tapping, the
technical literature indicates that a cutting fluid can provide lubricating effects
that serve to
reduce friction levels, and avoid such undesirable phenomena as chip clogging.
Cooling and Heat Transfer
When a cutting fluid provides lubrication to a machining process, it serves to
reduce friction
levels and thus moderate increases in temperature (Merchant, 1958). For many
machining
operations, however, the principal role of the cutting fluid is to remove heat
during the process,
especially from the zones indicated in Figure 1. This is especially true for many
modern-day
processes, in which potential lubricity benefits of a cutting fluid do not occur
because of the high
speeds utilized during the process. The success of the cutting fluid in providing
cooling can be
measured in different ways. A major focus has been on defining the temperature and
its
distribution in the cutting zone. Early attempts to characterize temperature were
done through
experimental methods. They used a variety of techniques such as: i) thermocouple
methods (e.g.,
tool-work thermocouple and embedded thermocouple) (Shore, 1925; Gottwein, 1925;
Herbert,
1926; DeVries, 1968, Watanabe et al., 1977; Hirao, 1989; Agapiou and Stephenson,
1994), ii)
infrared imaging (Schwerd, 1933; Boothroyd, 1961), and iii) microstructural change
(Trent,
1984).
Fig. 1 � Regions of heat generation in machining
Since experimental methods generally provide only limited information on complete
temperature
distribution, researchers have also employed models to establish temperature
distributions. Shaw
et al. (1951) studied the resistance across the tool-chip interface and related
this to temperature.
Chao and Trigger (1955) established models to predict the temperature on/near the
shear plane
using the assumption that all of the mechanical energy associated with shearing was
converted
into thermal energy, an assumption that is widely used in metal cutting. Childs et
al. (1988)
compared finite element analysis results to experimental data in predicting the
effects that a
cutting fluid has on the temperature distribution in machining. This study found
that the
effectiveness of heat removal was dependent on the flow rate and application
direction of the
fluid. Subramani et al. (1993) established a three-dimensional analytical model for
predicting
the workpiece temperature distribution during dry cylinder boring. The model was
verified by
measuring the temperature in cast iron cylinder boring tests. Zheng et al. (2000)
developed a
model to predict the temperature distribution in the workpiece during cylindrical
boring of cast
aluminum alloys under wet conditions.
Previous research has examined the perceived benefits of cutting fluids across a
wide variety of
cases. It has also cast some doubts on the necessity of cutting fluid use in some
machining
processes and under certain conditions. The most notable conclusion of the studies
is that more
knowledge is required to further quantify the role that cutting fluids play.
Furthermore, greater
information is needed to understand how cutting fluid mist, the source of many
potential health
concerns, is produced and behaves.
Cutting Fluids & Air Quality
The application of a cutting fluid stream to a rotating cylindrical workpiece, such
as found in a
turning operation, is illustrated in Fig. 2. As is evident, the application of the
fluid produces
airborne particles of varying sizes. Cutting fluid mist (especially the small
particulate that can be
inhaled and is too small to be seen in the figure) produced during machining
operations may pose
a significant threat to worker health/safety. Safety/health regulations focus on
the time weighted
average of the mass concentration of fluid mist to which a worker may be exposed
for a given
work period. Common strategies to control the amount of mist exposure include the
use of
enclosures, air filters, and mist collectors. However, these approaches prove to be
costly both in
time, as access to machine tools may be restricted by enclosures thus increasing
part
loading/unloading time, and money, including the slowing of production and
reduction in
process efficiency (Leith et al., 1996). In order to formulate a strategy to combat
this problem,
there must be an understanding of i) air quality and its harmful effects and the
impact of mist on
air quality, ii) the mist formation process, and iii) mist behavior.
Fig. 2 � Cutting fluid applied to a rotating cylindrical workpiece
Air Quality and Fluid Mist Basics
Several studies have shown statistically significant increases in cancer of the
esophagus, stomach,
pancreas, colon, prostate, and rectum due to prolonged exposure to cutting fluid
mist (Hands et
al., 1996; Mackerer, 1989). Because of such health problems a number of government
agencies
such as the National Institute of Occupational Safety and Health (NIOSH), the
Environmental
Protection Agency (EPA), and the Occupational Safety and Health Administration
(OSHA) have
become involved in establishing standards and regulations for particulate exposure.
Standards
set by industrial organizations and government agencies closely follow the U.S.
National
Ambient Air Quality Standards (NAAQS) established by the EPA. In 1987, these
standards set a
maximum mass concentration for PM10, particulate matter less than 10 micrometers
(also
referred to as thoracic particulate mass). This subset of particles represents the
portion of
inhalable particles that pass the larynx and penetrate into the conducting airways
and bronchial
regions of the lungs. Larger particles that enter this region can be evacuated from
the body in a
short amount of time. In 1997, in response to a growing concern about smaller
particles posing a
greater risk to human health, the standard was modified to address the risk
associated with
inhaling particles less than 2.5 micrometers in diameter (PM2.5). PM2.5 represents
the fraction
of respirable particles that enter into the deepest part of the lungs, the non-
ciliated alveoli.
NIOSH recommends that exposure to thoracic particulate mass be limited to 0.4 mg/m3
as a time
weighted average concentration for up to 10 hr/day for a 40 hr work week (NIOSH,
1984).
In an effort to understand the dominant factors in the production of potentially
harmful cutting
fluid mist, much work has focused on identifying statistically significant
variables of the process.
For example, Gunter and Sutherland (1999) investigated the application of fluid to
a rotating
cylindrical workpiece. This study found that the most significant variables in the
production of
PM10 and PM2.5, respectively, were the spindle speed and workpiece diameter (both
contributing to surface velocity). A study by Yue et al. (1996) reported similar
findings, noting
that spindle speed plays the dominant role in influencing aerosol mass
concentration for both
PM10 and PM2.5 mass concentration levels, and that significant submicrometer mist
particles
were produced in the process. Increasing the spindle speed appeared to increase the
amount of
submicrometer particles produced. Sutherland et al. (2000), in a study comparing
wet and dry
machining, found that a significant amount of submicrometer size fluid mist
particles were
produced during the turning process, and increasing spindle speed increased the
quantity of these
particles. The study also found that feed rate and depth of cut were significant
factors in the
formation of both cast iron dust and fluid mist. They also contrasted air quality
conditions,
quantified by mass concentration, under wet and dry machining conditions. A
randomized 23
factorial experiment with one replication was designed and conducted, varying
cutting speed,
feed, and depth of cut. The experimental set was run once in the presence of a
cutting fluid (5%
synthetic solution), as well as once in the absence of the fluid, yielding thirty-
two total tests.
Mass concentration was used as the response variable in an analysis of variance,
and all main
effects were found to be significant in generating mist and dust. The amount of
aerosol (mist for
the wet tests and dust for the dry tests) generated in the tests was found to be 12
to 80 times
greater in wet machining than in dry machining.
To prevent or counteract the problems associated with the cutting fluid mist
produced in
machining operations, there first must be an understanding of the mist formation
process. As
illustrated in Fig. 3, two different mechanisms have been proposed as sources for
cutting fluid
mist: atomization and vaporization/condensation. Both mechanisms of mist formation
will be
discussed and their significance explored.
Fig. 3 � Cutting fluid mist generation mechanisms
Cutting Fluid Mist Formation
Some recent work has been focused on examining the formation and behavior of
cutting fluid
mist. A predictive model has been established, with specific emphasis having been
placed on
mist formation in the turning process. A summary of this model development and its
validation
follows.
Atomization, one of the mechanisms by which cutting fluid mist is produced, is the
process by
which a liquid jet or sheet disintegrates by the kinetic energy of the liquid
itself, by exposure to
high-velocity air, or as a result of mechanical energy applied externally through a
rotating or
vibrating device (Bayvel and Orzechowski, 1993). In the last half century, there
have been many
studies conducted on liquid atomization due to interaction with a rotating element.
Most
attempts to model the process have characterized the rotating element as a disk, as
shown in Fig.
4. A liquid stream with a flow rate of Q is added at the center of the rotating
disk. Carried by
friction, the liquid spreads as a film towards the outer edges. At the rim of the
disk, the liquid
disintegrates into droplets. The liquid disintegration process is dependent on many
variables.
Depending on characteristics like flow rate, angular speed, disk diameter,
viscosity, and surface
tension of the fluid, it has been suggested that the fluid will disintegrate via
three different modes:
drop mode, ligament mode, and film mode. These modes are depicted in Fig. 5.
Fig. 4 � Fluid application to a rotating disk
Of special interest in characterizing the atomization process for a rotating disk
are the mean
diameter of the droplets produced by the different disintegration modes and also
the transition
stages between modes. The drop mode occurs when flow rate is fairly low. Matsumoto
et al.
(1985) predicted the mean diameter of droplets formed in drop mode as follows:
D We R d
= 3.2 -0.523 , (1)
where R is the radius of the disk and We is the Weber number. Based on the Weber
analysis
(1931) for the instability of a Newtonian jet, the mean droplet diameter formed
from the ligament
mode is approximated as follows:
L lig D = 1.88d , (2)
where dlig is the diameter of the ligament. Matsumoto et al. (1985) also reported
that the mean
diameter of the droplets formed in film mode could be predicted by the following:
11.1 Re 0.5 ( 0.1 QF0.2 )
F D = R � - We - + , (3)
where Re is the Reynolds number and QF is the volumetric flow rate to form the
film.
Fig. 5 � Three modes of atomization: (a) drop mode, (b) ligament mode, and (c) film
mode
Making the assumption that transition between drop and ligament modes occurs as the
drop
departure frequency from the disk edge exceeds a certain value, Matsumoto and
Takashima
(1978) developed the following semi-empirical relation to predict the flow rate at
which
transition occurs:
Q = 0.096� 2 R2 v � �Re0.95 �We-1.15 LD p ? , (4)
where, QLD is the critical flow rate when the transition occurs. Matsumoto and
Takashima also
developed an expression for the transition between film and ligament mode, which
they assumed
to occur when the surface tension of the liquid was overcome by the inertial and
centrifugal
forces beyond the edge of the disk. In order to define the transition between film
and ligament
modes, QFL, the flow rate at this transition, is defined as:
Q = 0.34 � 2 R2 v � �Re0.667 �We -0.883 FL p ? . (5)
To better describe the behavior of the cutting fluid involved in the turning
process, atomization is
characterized using a cylindrical workpiece. Given the geometry of the turning
process, it is
clear that the flow rates in the rotating disk theory are different from those in
the rotating
cylinder theory. This means that the transitional flow rates defined in Eqs. 4 and
5 are not valid
for these conditions. In the case of a cylindrical workpiece, two rims of fluid
form on the surface
of the cylinder when the liquid is applied. While these rims behave similar to
synchronous
rotating disks, they contain the potential for all three mechanisms of mist
formation to occur
simultaneously as the flow rate changes around the circumference of the cylinder.
This
significant difference requires the introduction of a new term, flow flux, to
account for the
difference in conditions.
Flow flux is defined as the quantity of fluid purged from a unit length of the
circumference
during one time unit.
R
Q
F AB
?
=
, (6)
where, QAB is the flow rate of the liquid atomized from an arc (AB) on the surface
of the cylinder
and ? is the angle formed between points A, B, and the centerpoint of the cylinder.
Fig. 6
illustrates the different mist formation mechanisms and the flow rates
corresponding to those
regions of the cross section.
Fig. 6 � Droplet formation and corresponding flow rates on a cylindrical workpiece
Making critical assumptions regarding the fluid flow based upon the visual evidence
of the
modes shown in Fig. 6, the mean droplet size and particulate distribution (shown in
Fig. 7) can
be predicted for each of the three formation mechanisms. These relations are
D We R D
= 3.2 -0.523 , (7)
( ) 7
2
3 2
7
1
3
2
7
2
1 23 . 1 ? ?
?
?
? ??
?
? ?
?
?
? ?
?
?
=
? ?
s
s
?
R R
Q
D R L
L Nl
, (8)
11.1 Re 0.5 ( 0.1 QF0.2 )
F D = R � - We - + , (9)
where ? is the cutting fluid density, QL is the volumetric flow rate to form
ligaments, s is the
cutting fluid surface tension, ? is the angular velocity, and QF is the volumetric
flow rate to form
film. The number of ligaments, Nl, is determined from the work reported by Kayano
and
Kamiya (1978) that concentrated on rotating disk atomization:
( ) ,
1
2 3 8 3 1 1
? ??
?
? ??
?
? ?
?
?
? ?
?
?
= + - + +
N St
We N N St
l
l l (10)
where St is the stability number.
Fig. 7 � Particulate Size Distribution due to formation and behavior
Runoff is a phenomenon that can occur in typical machining settings. Runoff usually
occurs in
the film mode region when the fluid flow rate is very high. The fluid in the film
region becomes
very unstable and drains off. Despite this occurrence, the drop and ligament modes
that are
occurring are not affected by it and the model still holds for those modes.
Vaporization of cutting fluid under high temperatures, the second mist production
mechanism,
could present a significant problem due to the organic compounds contained in the
fluid that
have been suggested to pose negative health effects. The fluid may vaporize or even
boil during
the machining process since the temperature in regions well away from the cutting
edge may be
in excess of 200�C; surface temperatures close to the cutting edge can be as high
as 650�C
(Childs et al., 1988). Concerns related to the relative importance of vaporization
(and
subsequent condensation of fluid vapors to produce fluid mist) prompted a study to
determine the
significance of vaporization in addressing concerns with cutting fluid use.
A 10% water diluted Chempakool 3121 concentrated cutting oil was examined under
high
temperature and the vapor generated was condensed into liquid and analyzed through
a
gravimetric analysis and a Gas Chromatography/Mass Spectrometry analysis. The
results of the
analysis showed that approximately 99.87% of the cutting fluid vapor that was
produced was
water vapor, thus not representing a health concern.
Cutting Fluid Mist Behavior
Having developed a model for cutting mist formation, an explanation of behavior
will complete
the characterization. Once cutting fluid mist has formed, it undergoes three
processes: settling,
evaporation, and diffusion. Settling is typically the result of two forces: gravity
and resistance to
droplet motion by the surrounding gas. Stokes� Law is often used to characterize
the settling
velocity. However, since much of the focus on cutting fluid mist is on particles
produced below
10�m in diameter, and these particles tend to fall faster than others because of a
�slip effect� at
the surface of the particle, an adjustment must be made to Stokes� Law as follows:
c
d
s gC
d
V
?
?
18
� 2
= , (11)
Where ?d is the cutting fluid density of a particular droplet, d is the diameter of
a particular
droplet, g is the acceleration due to gravity,
??
?
??
?
??
?
??
?

= + � + - -
-
8
8
6.6 10
2.514 0.8exp 0.55
6.6 10
1
d
d
Cc is the empirical slip correction factor, and
? is the small radial position within the fluid film.
Experiments have demonstrated that Eq. 11 extends the range of Stokes� law to
particles below
0.1�m in diameter (which corresponds to a Reynolds number less than 1.0) (Hinds,
1982).
As a droplet is settling in the air, both evaporation and settling mechanisms are
occurring
simultaneously and are influencing each other. The size of the droplet decreases
due to the
evaporation, which leads to a smaller settling velocity. In turn, the slower
relative motion of the
droplet will decrease the evaporation rate. The transient velocity of each droplet
can be
calculated by Eqs. 11 and 12, while the transient mass is defined as:
M
G
H
m
Sc
Sh
dt
dm
??
?
??
= - ? ?
?
?
??
?
3 t
, (13)
where Sh is the Sherwood number, ScG is the Schmidt number of a surrounding gas, m
is the
mass of the droplet, and HM is the specific driving potential for mass transfer.
Also, the transient
temperature is given by:
( )
( ) m
m
C
L
e
Nu T T
dt
dT
L
G v
G
&
? ??
?
? ??
?
+
-
-
???
??
? = ?
?
?
??
?
3Pr 1
1


t
?
, (14)
where Nu is the Nusselt number, Pr is the Prandtl number, ? is the ratio of
surrounding gas heat
capacity to that of the liquid phase, Lv is the latent heat of evaporation, CL is
the heat capacity of
the liquid, and m& = dm/ dt is negative for evaporation. The subscripts on the
variables denote
the gas phase away from the droplet surface (G), the vapor phase of the evaporate
(v), and the
liquid phase (l). The particle time constant for Stokes� flow is d G t = ? d 2 /18�
, and � is the nondimensional
evaporation parameter.
Cutting fluids are typically mixtures; the fluid composition examined by Yue et al.
(2004) and
Sun et al. (2004) was 95% water combined with 5% soluble oil-based cutting fluid
concentrate.
Yue et al. (2004) and Sun et al. (2004) established a mist behavior model that
considered settling
and evaporation, and used this to predict the change in particle size within a work
volume over
time, and also the particulate mass concentration within the volume. The validation
data for
mass concentration, depicted in Fig. 8, shows a good agreement between the
predicted behavior
and the actual data.
Fig. 8 � Measured and predicted PM10 during settling
and evaporation [Yue et al., 2004; Sun et al., 2004]
Novel Approaches to Controlling Cutting Fluid Mist
In some operations, eliminating or reducing the amount of cutting fluid used is
thought to be
practically impossible. In order to minimize health concerns, it may be
advantageous to utilize
the knowledge of the operation to employ cutting fluid mist control strategies. The
following
section discusses some of these strategies.
Kinematic Coagulation
A novel approach for eliminating the cutting fluid mist is to capture the small
airborne particles
with larger fluid droplets. The strategy of kinematic coagulation, the capture of
smaller particles
by larger collector droplets, was explored by Kinare et al. (2004). These studies
observed the
effects of the strategy on the airborne mist produced during a turning process.
After a field of
mist particles has been created by the application of a cutting fluid during a
turning process, an
atomizer then sprayed collector droplets of a controlled size into the field of
motion of the
process-generated mist particles. The particles from the atomizer collide with the
other particles
and coalesce (Fig. 9), which created large particles that rapidly settled.
Fig. 9 � Illustration of kinematic coagulation strategy
The behavior of the collector droplets is of particular importance while attempting
to reduce the
number of small particles in the air. When collision between a process generated
particle and a
collector droplet occurs, four possible behaviors can result. The first possibility
is that the two
particles could bounce off each other. They could also collide and permanently
coalesce.
Particle disruption could also occur, in which the particles temporarily coalesce,
then break apart
again. Particle fragmentation could also possibly occur in the event that one
particle rips past the
other and pulls some smaller particles with it. Of course, since it is desired to
remove the
process generated mist particles from the breathable airspace, the goal of the
kinematic
coagulation work is to identify conditions that promote collision and permanent
coalescence.
Previous research observed that permanent coalescence is promoted when the ratio of
colliding
droplet diameters was between ten and fifteen (Kinare et al., 2004). Smoluchowski
(1917)
provided a relation that allows for estimating the number of droplets that need to
be introduced in
order to accomplish the desired kinematic coagulation of the mist particles. A
custom designed
atomizer system was developed to produce the desired size (100-200 micrometer) and
number of
collector droplets to interact with those produced by the turning process. To
assist with the
atomizer design, a Malvern Series 2600 Particle Sizer was used to measure the
droplet size
distribution produced by the atomizer. Kinare et al. (2004) noted that
coagulation/coalescence
will only occur to a significant degree for particle concentrations in excess of
1012/m3, and that
collisions will occur more frequently as the concentration is increased. Air
quality data collected
during a turning operation indicated particle concentrations in the range of 109-
1015/m3, and this
observation was used to define additional atomizer characteristics.
Once the atomizer was designed, experiments were performed to assess its efficacy
(Kinare et al.,
2004). The addition of the atomizer droplets had a significant impact on the amount
of fine
particles present in the air. Data from the Particle Sizer indicated that the
atomizer was able to
reduce mist levels by approximately 20%. These experiments concluded that of the
variables
present within the process, the use of the atomizer was the most relevant to the
mass
concentration, count mean diameter (CMD), and number of particles. Furthermore,
when
atomizer droplets were present, mass concentration depended heavily upon pump
pressure and
nozzle geometry of the atomizer system. Most importantly, the work demonstrated the
ability to
reduce airborne particulate matter through the application of the kinematic
coagulation
mechanism.
Innovative Machine Tool Design
In spite of advances in machining technology, there will undoubtedly remain some
processes that
will have to be performed wet in order to achieve the desired production rate and
part quality.
For these processes, machine enclosures are widely employed to control air quality
by containing
the mist and thereby protecting worker health. However, construction of an
enclosure around the
machine tool serves to only temporarily contain the machining generated mist.
Machining mist
can escape, possibly in high mass concentrations, from openings in the enclosure.
Also, opening
the enclosure access door upon completion of a machining process allows accumulated
machining mist to enter the workers� breathing zone. The effectiveness of a machine
enclosure
is directly related to its ability to contain the cutting fluid mist produced
during a wet machining
process to prevent deterioration of the air quality in the workers� environment.
Intelligent
machining system design, which considers the movement of machining generated mist
subsequent to its generation, can make machine enclosures more effective.
Air quality in the workplace is affected not only by the amount of mist that is
generated during
the machining process, but also by its motion after it is created. While mist is
produced in the
vicinity of the machine tool and workpiece, there are several mechanisms that
disperse the mist
throughout the workspace. These mechanisms include air currents created by HVAC
systems,
mist exhaust systems, and/or the motion of the workpiece or the machine tool. In
the presence of
a machine enclosure (without a mist exhaust feature) the influence of the HVAC
system on mist
motion is eliminated and the movement of the mist inside the enclosure depends
primarily on the
movement of the machine tool or workpiece and the shape of the enclosed space. This
machining scenario can be numerically simulated to analyze the flow patterns that
are produced
and the distribution of the machining generated mist within the machining
enclosure.
To investigate the airflow effect on cutting fluid generated mist during a turning
process a
computational fluid dynamic (CFD) study was conducted (Hii, 2005). To simulate the
movement of the mist within the machine enclosure an accurate representation of the
enclosed
space was constructed, appropriate boundary conditions were applied to simulate the
motion of
the workpiece, and the resulting airflow was simulated before cutting fluid
droplets, representing
both ambient and machining generated mist, were incorporated into the model. Using
this model,
parametric studies were performed by varying the workpiece diameter and rotational
speed to
investigate the effect on mist motion within the enclosure.
The CFD model was constructed using Star-CD. The model geometry was a cross
sectional
plane of the experimental enclosure around a lathe and was representative of the
turning
experiments conducted by Kinare et al. (2004). The computational space is bounded
by the
machine enclosure and includes the workpiece, the tool holder, and the chip/drip
pan. Since the
cutting fluid mist is generated primarily at the cutting zone and in its immediate
vicinity, a twodimensional
model of the machine enclosure was deemed sufficient for this preliminary study.
The first part of the preliminary study consisted of simulating the airflow inside
the enclosure for
various combinations of workpiece diameter and rotational speed. The workpiece
diameter was
varied between 2 inches and 4 inches and the workpiece rotational speed was varied
between 400
rpm and 2000 rpm. All simulations were conducted for 4 minutes of process operation
with a
constant spindle speed. In all cases it was observed that the velocity field
achieved a steady state
condition after approximately 1 minute. There is a rotating air stream around the
workpiece and
a recirculation zone directly above the workpiece. Outside of these two regions the
airflow
velocity remains low.
Once the airflow velocity and pattern were obtained for the experimental enclosure
under
different workpiece diameters and workpiece rotational speeds, fine droplets were
introduced to
study the effect of airflow on the distribution of droplets and to observe any
potential interaction.
Fine droplets having a diameter of 1 micrometer were introduced directly below, to
the left, over
the top, and diagonally in the left upper quadrant of the model. Droplets
representing suspended
ambient mist were assigned initial locations slightly away from the workpiece
surface, while
machining process generated droplets were located in close proximity to the
workpiece surface.
In addition, machining process generated droplets were assigned an initial velocity
based on the
workpiece rotational speed and suspended ambient droplets were assigned an initial
velocity of
zero.
The air stream around the rotating workpiece effectively entrained the fine
droplets that were
within close proximity to the workpiece, as can be seen in Figure 10, which
displays some
representative results from the study. Fine droplets that were not entrained by the
rotating air
stream were carried by the recirculation zone air movement. The degree of
entrainment was
related to the workpiece tangential velocity. A larger workpiece diameter or a
higher workpiece
rotational speed resulted in a faster rotating air stream around the workpiece,
which directly
increases the level of droplet entrainment. Conversely, a smaller workpiece
diameter or a lower
workpiece rotational speed allows more fine droplets to be carried around the model
by the
recirculation zone air movement to become more evenly dispersed within the machine
enclosure.
It should be noted that the presence of fine droplets did not significantly affect
the airflow
magnitude and pattern.
Fig. 10 � CFD Model of Airflow and Mist Behavior in a Machine Tool Enclosure
This preliminary study points to a promising and exciting strategy for dealing with
machining
mist: design of the enclosure and machine tool elements. A CFD model such as that
described
above, can be used to assess the impact of changes in the enclosure geometry,
cross-slide
position and geometry, drip/chip pan position and geometry, etc. on air flow and
mist behavior.
Airflow inlets and exhaust ports may also be configured to control mist and thus
limit the amount
of worker exposure.
Wet versus Dry Machining
In previous sections, the technical literature has been reviewed to characterize
prior work that
has considered the role of cutting fluids in machining operations. Recent efforts
that have
focused on describing the formation and behavior of cutting fluid mist have also
been presented.
Evident from the discussions to this point, it may be desirable to greatly reduce
or even eliminate
cutting fluids from machining operations. Certainly, depending on the tolerances
associated with
the desired output of the machining process, different processes are more suitable
for reduction
or elimination of cutting fluid. Because of its accessible cutting zone, turning
operations have
the potential to be performed dry. Sawing and milling also provide excellent
opportunities to
eliminate cutting fluid use because their interrupted nature ensures short breaking
chips, good
chip clearance, and cooling of the cutting edges (Diniz and Micaroni, 2002). Hole-
making
processes like drilling and tapping are often hard to accomplish without some fluid
lubricant,
since chip removal is a key to process efficiency. With these thoughts in mind,
some recent
contributions focused on dry versus wet applications are provided below.
Open Faced Operations
An open faced machining operation is one where the interaction surfaces between the
tool and
the workpiece are readily accessible for easy chip removal and cutting fluid
application.
Examples of open faced operations are milling, sawing, cylinder boring, and
turning. These
operations are those that tend to utilize cutting fluids primarily for their
cooling and heat transfer
capabilities. Transferring the heat produced by a machining operation away from the
cutting
zone, tool, and workpiece may be important for at least two reasons: i) thermal
distortions can
lead to poor machined component accuracy, and ii) elevated temperatures can lead to
higher tool
wear rates. Therefore, in considering dry (or nearly dry) versus wet machining
scenarios, these
reasons have dominated recent discussions in the literature.
One of the challenges of dry machining is the concern about increased wear rates in
the absence
of a cutting fluid (Weinert et al., 2004), and researchers continue to explore
techniques to
attenuate these elevated wear rates, one of which is improved cutting tool
materials. Cemented
carbides are, by far, the most widely used tool material. As a rule of thumb, as
the grain size
becomes finer, the more wear resistant the material becomes (Dreyer et al., 1997).
Reducing the
grain size of tungsten-carbide powders below 0.8 �m equates to a set of conditions
in which even
small tools can be produced to have good cutting edge stability and dry machining
of high
strength materials is plausible (Byrne et al., 2003). Some applications require
characteristics
other than those provided by the cemented carbides at elevated temperatures. In
these cases
cermets can be used as cutting materials. Cermets have a higher hot hardness in
comparison to
cemented carbides, and thus make it possible to cut at higher speeds. Cermets also
have
excellent chemical stability against oxidation and tribochemical wear and a reduced
affinity for
diffusion, due to the ceramic component (Porat and Ber, 1990). To withstand
machining of
harder materials like gray cast iron and hardened steels, while enduring high
temperatures and
providing a longer tool life, ceramic materials are applied. This class of cutting
materials is
known for high hot hardness and reduced resistance to thermal shock (making it
possible to cut
without the aid of a coolant) but suffers from a lack of toughness. Turning and
milling tests
performed on tools with SiC reinforcements and a very fine aluminum powder by
Narutaki et al.
(1997) indicate that the new material permits significantly higher limits on feed
rate and lower
wear rates. Like ceramics, cubic boron nitride (CBN) tools are generally used in
machining of
harder materials like cast iron and hardened steel because of its excellent high
hardness and
chemical wear resistance characteristics (T�nshoff et al., 1995). Polycrystalline
diamond (PCD)
is the hardest tool material available, and it is often used to machine light
metals like aluminum
and magnesium. Despite high strength, low coefficient of friction, and other
positive qualities,
its application is limited due to the graphitization phenomenon that occurs above
600�C (Byrne
et al., 2003). Tooling characteristics can play a large role in determining if a
process should be
run in wet or dry conditions.
Tool coatings can reduce the rate of abrasive and adhesive wear by acting as a
barrier between
the cutting tool substrate and the workpiece material. A wide variety of different
coating
materials and coating strategies exist. Multilayer coatings combine the favorable
characteristics
of each coating and distribute stress better than monolayer coatings. They also
relieve crack
energy by deflection and branching, thus delaying tool failure. Technological
advancements
have recently allowed for the use of coatings on the order of nanometers (Ducros et
al., 2003;
Cselle et al., 2003). Since more layers are desirable and the cutting edge radius
is dependent on
the total thickness of the coating, it is easy to see the advantages of these very
thin coatings. A
high performance class of coatings called supernitrides has also been shown to
exhibit improved
wear behavior in a set of dry milling tests on 42CrMo4V (Erkens et al., 2003). The
use of very
hard coatings like CBN for machining ferrous materials and CVD (chemical vapor
deposition)
diamond for machining of non-ferrous metal alloys helps to reduce wear given a
demanding set
of circumstances.
One of the oft-quoted benefits of a cutting fluid in an open faced machining
operation is its
ability to transfer heat. Heat transfer can be of substantial benefit in the
reduction of surface
error, a measure of the deviation of the machined surface from that of a surface
produced under
ideal conditions. In particular, a lack of poor machined cylindricity in engine
bores can produce
poor performance due to increased oil consumption, frictional loss, and excessive
wear of piston
rings. With this in mind, Cozzens et al. (1995) examined the difference between wet
and dry
machining of Al 308 and Al 390 die cast cylinders. Temperature along the length of
the bore
(measured by thermocouple probes positioned along the workpiece axis), forces, and
coolant
temperature were measured for each test. In the presence of a cutting fluid, the
temperature
initially spiked at each position, but at a significantly smaller magnitude than
during dry cutting.
The peak and average temperature was reduced by as much as 50% when comparing the
wet and
dry machining tests.
In another study of boring 308 die cast aluminum cylinders on a vertical milling
machine, Zheng
et al. (2000) studied the effects of cutting fluid use via inverse heat transfer
and finite element
methods to determine estimates of the effective convection coefficient that the
fluid provided.
The purpose of the work was to determine the effect that cutting fluid presence had
on
temperature and surface error produced by the process. The study also successfully
developed
an analytical model to predict temperature distribution, and observations regarding
performance
measures of the process were made. Fig. 11 depicts the two dimensional governing
equation of
the heat transfer in a cylindrical bore. The equation, which considers a
longitudinally moving
heat source and heat losses through the inside and outside of the bore walls, is
t
w
k
g z t
H
dz
d
w
?
- + = ??
a
? ? ( , )
2 2
2
, (15)
where, ? is the difference between the wall and ambient temperatures, a is the
thermal diffusivity
of the material, k is the thermal conductivity, H is the ratio of heat convection
coefficient to
thermal conductivity (h/k), and g(z,t) is the heat source strength. Equation 15 is
subject to the
following boundary and initial conditions:
+ = 0
?
?? ?
H
z
at z = 0, and L respectively
? = 0 at t = 0, for all z. (16)
where, the heat source strength at cutting tool position zs in the boring operation
may be
estimated by:
( , ) ( ) s s s g z t = g �d z - z . (17)
Fig. 11 � Two dimensional heat transfer in a cylinder bore
The term zs may be further expressed in relation to feed and the initial position
of the tool. The
proposed model was consistent with the measured data. Furthermore, it was clear
that the
thermal expansion, rather than the elastic deflection from the cutting forces, was
the dominant
factor in influencing surface error. Also, the peak magnitude of the surface error
was smaller in
the presence of a cutting fluid. The introduction of a cutting fluid reduced the
surface error from
that produced in dry boring by approximately one half.
In an effort to better understand the role of cutting fluids as coolants, Shen et
al. (2001)
developed an analytical model for predicting the workpiece temperature in
peripheral milling.
First, a set of dry milling tests were performed. Assuming heat transfer by means
of natural
convection during dry milling and using an inverse heat transfer method, the heat
source strength
was calculated. Next, using the cutting power data, the fraction of heat
transferred to the
workpiece in the operation could be estimated. Then, the heat fraction for the set
of wet milling
tests is assumed to be the same as the counterpart in the dry tests. This allows
for the heat source
strength to be estimated again, and now using the cutting power and the heat
fraction data, the
convection coefficient of the cutting fluid can be estimated based on the
temperature data. Table
1 depicts some of the quantitative estimates of the cutting fluid convection
coefficients.
Table 1 � Estimated convection coefficients in peripheral
milling (shaded tests employed a cutting fluid)
These investigations yielded some important conclusions. Increases in the speed,
feed, and depth
of cut yield an increase in the heat source strength. Heat transferred to the
workpiece also
increases. Furthermore, the estimated convection coefficient for wet milling was
approximately
three orders of magnitude larger than that during dry milling, indicating the
ability of the fluid to
function as a coolant during this process.
Closed Faced Operations
Closed face operations are those that have a fairly inaccessible cutting zone and
interaction
surfaces that are not easily accessible for chip removal and cutting fluid
application. Examples
of closed face operations are drilling, tapping, and reaming. These are the
operations that
primarily utilize cutting fluids for their lubricating properties. Studies on both
drilling and
tapping indicate that the presence of a cutting fluid impacts the machining forces
or spindle
torque required for these operations. These studies also established that the
presence of a cutting
fluid significantly improves surface finish.
Haan et al., (1997) performed a series of eight test plans to measure a group of
drilling outputs,
including torque and thrust force, as well as performance outputs like surface
finish and hole
quality, to determine the most significant effects given different configurations.
Standard
factorial design methods were employed, and normal probability plots were used to
determine
statistically significant effects. Presence of a cutting fluid did affect the
torque data. In all but
one test plan, the average torque was reduced when a cutting fluid was present. Two
key points
should be noted. First, feed was consistently found to be a more significant
variable than the
presence of a cutting fluid. Secondly, no significant effect was found when
comparing results
between the use of 2% and 8% concentration cutting fluid (water soluble oil). An
even more
significant observation in the drilling experiments was the effect the cutting
fluid had on surface
finish and hole quality results. By inspection, it was quite clear that the hole
created without the
aid of a cutting fluid had a significantly larger average surface finish, as well
as a significantly
larger variation.
Cao and Sutherland (2002) found that similar improvements were made in the tapping
process
when cutting fluid was present. Initial tests found that there was considerable
inconsistency in
tapping torques and axial forces even under the same machining conditions. The
torque data was
found to consist of two parts. The first part is the base load which constituted
the contribution
due to chip formation and overall tool/workpiece friction. They established a model
for the
process that considered this effect, with the friction behavior dependent on
measured fluid-based
properties. The second part is termed the chip packing load, and is due to the
chips clogging the
flutes of the tap and causing excessive torque.
From Fig. 12, it appears that the model fairly accurately predicts the base load.
Since the chip
clogging phenomenon was not taken into account in this model, the random loading
spikes are
not predicted by the model. While a tapping oil provided a noticeable reduction in
torque and
axial force, soluble and straight oils either provided no benefit to the operation
or actually
increased the torque and force.
Fig. 12 � Modeled and measured tapping loads due to
different lubrication conditions
The tapping experiments revealed that the use of cutting fluid can reduce the
friction in the
process (assuming the appropriate fluid is selected). The coefficient of friction
significantly
changes in the presence of a cutting fluid; friction was found to be approximately
four times
larger for dry cutting as opposed to wet cutting. The experiments also showed that
chip packing
load was a significant factor in terms of such measures as thread quality and tap
breakage.
In a study of cutting fluid conditions in the boring of die cast aluminum alloy
tubes, the presence
of a fluid, regardless of concentration, did not significantly affect the machining
forces (Cozzens
et al., 1999). Furthermore, normal probability plots showed that the cutting fluid
had no effect
on built up edge (BUE) or the surface finish of the machined part. This study
further lends
support to the assumption that the use of a cutting fluid only plays a significant
role in machining
operations where the primary need is lubricity.
In the absence of a cutting fluid to provide lubrication in closed face operations,
other methods
must be explored to provide this function. One promising dry machining enabling
technology is
advanced cutting tool coatings. As noted previously, coatings can retard the
process of tool wear;
but, they can also be employed to enhance tribological behavior. Some soft coatings
are
considered self-lubricating, which reduce the friction between the tool and the
workpiece and
therefore reduce the cutting forces and heat generated (Byrne et al., 2003;
Derflinger et al., 1999).
Novel Approaches to Reduction of Cutting Fluid Use
As discussed, eliminating or reducing the amount of cutting fluid in some processes
could be
done without compromising performance measures. In these cases, operations can be
performed
to satisfaction in a more economically feasible and environmentally responsible
manner. The
following section discusses some of these strategies.
Minimum Quantity Lubrication
Minimum quantity lubrication (MQL) is a strategy that can offer technological and
economic
advantages over traditional fluid applications (Weinert et al., 2004; Klocke et
al., 1996). As the
name implies, MQL seeks to reduce the amount of cutting fluid used in an operation.
In terms of
technological advancement, MQL is to dry machining as a hybrid car is to vehicle
powertrains -
a step in the right direction. An MQL process can be performed with or without a
transport
medium, such as air, and a pump supplies the tool with the fluid - generally
straight oil - as a
rapid succession of precisely metered droplets. Quantitatively, MQL is associated
with the use
of between 10 and 50 mL of cutting fluid per machine hour. Emulsions and water are
usually
only used when it is essential to cool the tool more efficiently than is possible
with straight oil.
In contrast to the lubricating function, minimum quantity cooling (MQC) has been
largely
unexplored, but offers promise in some situations (Klocke et al., 2003).
Fig. 13 � Various MQL Systems [Weinert et al., 2004]
MQL Application
The method by which a fluid is added to the machining system, especially under
minimum
quantity lubrication conditions, can greatly affect the efficiency with which
cutting fluid
functions are performed. As illustrated in Fig. 13, the fluid can be applied in two
manners:
externally, through the use of separately secured nozzles, and internally, through
channels built
into the tool. Each application method has advantages and disadvantages (Klocke et
al., 2003;
Karino, 2002). A short discussion will highlight key features of both.
The external supply method is used in sawing, end and face milling, and turning
operations. In
the drilling, reaming, and tapping operations, this method is only appropriate when
the ratio of
l/d is less than three. If the operation parameters exceed that limit, then the
tool may have to be
withdrawn several times to be wetted again. Problems also exist in external supply
when more
than one tool must be used (Suzuki, 2002).
Internal supply is advantageous in the operations of drilling, reaming, and
tapping, where l/d
ratios tend to be large. This ensures that the cutting fluid is constantly
available close to the
cutting edge, and eliminates the concerns regarding nozzle positioning errors and
geometric
clearance issues inherent to supply pipes and nozzles. There are one- and two-
channel supply
systems available. In the two-channel system air and oil are fed separately through
the spindle,
and then are combined to form an aerosol just ahead of the cutting edge. Each
supply system has
limits on the amount of fluid it can supply to the cutting zone; system selection
will depend
heavily on process needs.
MQL Fluid Type
Cutting fluid selection is also a major consideration when evaluating the entire
machining system
(Weinert et al., 2004; Suda et al., 2002). Typically fluids are selected based on
their ability to
influence performance, as reported in some of the previously mentioned studies. Due
to low
consumption rates in MQL operations, secondary characteristics such as
biodegradability,
oxidation stability, and storage stability are more important because of
environmental
compatibility and chemical stability concerns.
Environmental compatibility is most heavily dependent upon biodegradability (McCabe
and
Ostaraff, 2001). Because of their advantageous biodegradability characteristics,
vegetable oils
have typically been used in MQL machining, while synthetic and polyol esters are
starting to be
considered more frequently. Storage is also a major consideration. Lubricant used
in an MQL
system must remain stable for long periods of time and under high temperatures.
MQL fluids fall into two primary groups: synthetic esters and fatty alcohols.
Synthetic esters
(generally vegetable oils) are commonly used because of their good lubrication
properties,
resistance to corrosion, high flash, and boiling points. However, fatty alcohols do
achieve better
heat removal and when vaporized, produce little in terms of residue as compared to
synthetic
esters. Synthetic esters generally are used in operations where lubrication is the
primary need for
a cutting fluid, whereas fatty alcohols are used in MQL applications that require
the cutting fluid
for heat removal (N.N., 1996; Suda et al., 2001).
MQL Performance
A number of MQL performance studies have begun to appear in the literature, and
Weinert et al.
(2004) discuss a number of these. Peripheral milling tests were recently performed
to examine
the effects of fluid application strategy (dry, MQL, and fluid flood), axial depth
of cut, flow rate,
and air pressure (Ju et al., 2005). The measured responses from these experiments
were cutting
force, workpiece temperature, machined surface error, and air quality. A synthetic
fluid, mixed
at a 5% concentration with water, was used in the tests in which cutting fluid was
applied. The
experiments in which the MQL fluid application was employed used an external
application
system in which the fluid and air were mixed in the process before being projected
through the
nozzle and to the cutting zone.
The work of Ju et al. (2005) concluded that while MQL application was not as
successful as
flood application in reducing the workpiece temperature, it did provide a sizeable
improvement
over dry machining. On the other hand, the measured forces were nearly identical
across the
three �fluid� strategies: flood, MQL, and dry. Not surprisingly, increases in fluid
flow rate and
air pressure were found to reduce temperature and improve surface finish.
Modulated Drilling
Another machining application that has seen an increased concentration of effort
for
environmentally responsible manufacturing is drilling. As discussed earlier,
drilling is one of the
few machining operations that require a cutting fluid mainly for its lubricating
properties. The
fluid is required in typical drilling operations to provide lubricity between the
chip and the tool.
Chip clogging is often the cause of drilling process failure, and given the nature
of drill flutes
and the continuous behavior of chips in this process, dry drilling has some
significant obstacles
to overcome.
Some success has been reported with an innovative dry drilling approach (Ackroyd et
al., 1998;
Toews et al., 1998). The work employed low frequency axial modulations of the
spindle
produced by a linear motor. This movement interrupted the built-up edge (BUE)
formation in an
attempt to achieve dry drilling. This method is credited with the ability to
control the chip
shape/size and produce chips that can be easily channeled out of the drill flute.
These smaller
chips may pass through the flutes much easier than the long chips that typically
cause adhesion
and clogging. McCabe (2002) reported on work to employ a similar method using high
frequency modulations instead of low frequency modulations. The limited success of
this effort
may have been due to the inability of the piezo-based actuator to produce high
enough oscillation
amplitudes.
The work of Filipovic and Sutherland (2005) concentrated on addressing the
limitations that
were reported by Toews et al. (1998) and McCabe (2002). A new magnetostrictive-
actuator and
tool holder assembly was used, as shown in Fig. 14, to initiate high frequency
modulations
during the drilling process. As the magnetic field is rapidly adjusted, the
magnetostrictive drive
rod elastically deforms to produce oscillations. This sinusoidal motion of the tool
results in the
production of discrete chips, especially for oscillation frequencies in the range
of 25-50 Hz.
Fig. 14 � Modulated drilling tool holder assembly and setup
To verify the improved performance characteristics of the process due to the
inclusion of the new
assembly, a 24-1 factorial experiment was conducted. Four variables of interest
were examined:
spindle speed, feed, frequency ratio (oscillation frequency/spindle rotation
frequency), and
workpiece material. As well as the beneficial effects observed due to modulation of
chip size,
this experiment clearly demonstrated that the axial drill oscillation can reduce
the process torque
and thrust force. The initial success of tool holder/actuator assembly in providing
axial drill
modulations appears to offer promise for the dry drilling of aluminum.
Summary and Conclusions
An extensive examination of the functions of cutting fluids and interactions with
other
machining system components has led to a general understanding of the roles that
cutting fluids
play in a metal cutting process. A model for the formation and behavior of cutting
fluid mist has
been presented. Health issues associated with exposure to cutting fluid mist have
been discussed.
Comparisons between wet and dry machining for several machining operations have
been made
using both analytical and experimental investigations, and findings have been
presented to
suggest that intelligent strategies may allow for the reduction or even the
elimination of cutting
fluids from certain machining processes. Novel approaches have been identified that
can be
utilized to eliminate or greatly reduce the amount of fluid that is needed for a
machining
operation. Novel approaches have also been proposed to eliminate or control cutting
fluid mist
for those situations in which, for the short term, a cutting fluid is deemed to be
a necessary
process requirement.
Acknowledgements
This work was supported, in part, by the Ford Motor Company, Chrysan Industries,
UNIST Inc.,
the NSF-ARPA Machine Tool Agile Manufacturing Research Institute (MT-AMRI), and the
National Science Foundation under Grant Nos. DMI-9502109, DMI-9628984, and DMI-
0070088. The authors gratefully acknowledge the research contributions of Steven
Batzer,
Tengyun Cao, Aleks Filipovic, Kenneth Gunter, Deborah Haan, Chuanxi Ju, Lucas
Keranen, Sid
Kinare, Walter Olson, Ge Shen, Jichao Sun, Yan Yue, and Yuliu Zheng in support of
this effort.
References
King, N.; Keranen, L.; Gunter, K.; Sutherland, J. Wet Versus Dry Turning: A
Comparison of
Machining Costs, Product Quality, and Aerosol Formation. SAE Paper, 2001, SP-1579.
Filipovic, A.; Olson, W.; Pandit, S.; Sutherland, J. Modeling of Cutting Fluid
System
Dynamics. Proc. of the 2000 Japan � U.S.A. Symposium on Flexible Automation. 2000,
Paper #2000JUSFA-13201.
Kalpakjian, S.; Schmid, S. Manufacturing Engineering and Technology; Prentice Hall,
Upper
Saddle River, NJ - USA, 2001, 585-590.
Iowa Waste Reduction Center (IWRC). Cutting Fluid Management for Small Machining
Operations. University of Northern Iowa. 1996.
National Institute for Occupational Safety and Health (NIOSH). Criteria for a
Recommended
Standard � Occupational Exposure to Metalworking Fluids. January 1998.
Mackerer, C. Health Effects of Oil Mists: A Brief Review. Toxicology and Industrial
Health.
1989, 5, 429-440.
Thorne, P.; DeKoster, J.; Subramanian, P. Environmental Assessment of Aerosols,
Bioaerosols, and Airborne Endotoxins in a Machining Plant. American Industrial
Hygiene Association Journal. 1996, 57 (12), 1163-1167.
Skerlos, S.; Rajagopalan, N.; DeVor, R.; Kapoor, S.; and Angspatt, V. Ingredient-
Wise Study
of Flux Characteristics in the Ceramic Membrane Filtration of Uncontaminated
Synthetic
Metalworking Fluids, Part 1: Experimental Investigation of Flux Decline. J. of
Manuf.
Sci. & Eng. 2000, 122 (4), 739-745.
Skerlos, S.; Rajagopalan, N.; DeVor, R.; Kapoor, S.; and Angspatt, V.
Microfiltration of
Polyoxyalkylene Metalworking Fluid Lubricant Additives Using Aluminum Oxide
Membranes. J. of Manuf. Sci. & Eng. 2001, 123 (4), 692-699.
Shaw, M. The Chemico-Physical Role of the Cutting Fluid. Metal Progress. 1942, 85-
88.
Ernst, H.; Merchant, M.; Chip Formation, Friction and High Quality Machined
Surfaces.
Surface Treatment of Metals. 1941, 29, 299.
Lee, E.; Shaffer, B. The Theory of Plasticity Applied to the Problem of Machining.
J. of Appl.
Mech. 1951, 18, 405-413.
Oxley, P.; Hastings, W. Predicting the Strain Rate in the Zone of Intense Shear in
which the
Chip is formed in Machining from the Dynamic Flow Stress Properties of the Work
Material and the Cutting Conditions. Proc. of Royal Society of London. Series A.
1977,
356 (1686), 395-410.
Iwata, K.; Osakada, K.; Teresaka, Y. Process Modeling of Orthogonal Cutting by the
Rigid-
Plastic Finite Element Method. Journal of Engineering Materials Technology. 1984,
41,
71-82.
Strenkowski, J.; Carroll, J. A Finite Element Model of Orthogonal Metal Cutting. J.
Eng. Ind.
(Trans. ASME). 1985, 107 (4), 349-354.
Lin, Z.; Pan, W. A Thermoplastic-plastic Large Deformation Model for Orthogonal
Cutting
with Tool Flank Wear � Part I: Computational Procedures; -Part II: Machining
Application. Int. J. Mech. Sci. 1993, 35(10), 829-850.
Marusich, T.; Ortiz, M. A Finite Element Study of Chip Formation in High Speed
Machining.
Manufacturing Science and Engineering. 1995, MED-Vo. 2-1/MH-Vol. 3-1.
Sabberwal, A. Chip Section and Cutting Force during the Milling Operation. Annals
of the
CIRP. 1960, 197-203.
Shaw, M.; Pigott, J.; Richardson, L. The Effect of Cutting Fluid upon Chip-Tool
Interface
Temperature. Trans. of ASME. 1951, 73, 45-56.
Cassin, C.; Boothroyd, G. Lubrication Action of Cutting Fluids. J. Mech. Eng. Sci.
1965, 7
(1), 67-81.
Haan, D.; Batzer, S.; Olson, W.; Sutherland, J. An experimental study of cutting
fluid effects
in drilling. Journal of Materials Processing Technology. 1997, 71, 305-313.
Cao, T.; Sutherland, J. Investigation of thread tapping load characteristics
through
mechanistics modeling and experimentation. Int. Journal of Machine Tools &
Manufacture. 2002, 42, 1527-1538.
Kahng, C.; Koegler, W. A Study of Chip Breaking During Twist Drilling. SME
Technical
Paper MR76-267. 1976.
Nakayama, K.; Ogawa, M. Basic Rules on the Form of Chips in Metal Cutting. Annals
of the
CIRP. 1978, 27 (1), 17-21.
Merchant, M. The Physical Chemistry of Cutting Fluid Action. Am. Chem. Soc. Div.
Petrol
Chem. 1958, 3 (4) A, 179-189.
Shore, H. Thermoelectric Measurement of Cutting Tool Temperature. Journal of
Washington
Academic Sciences. 1925, 15, 85-88.
Gottwein, K. Die Messung der Schneidentemperatur beim Abdrehen von Flusseisen.
Maschienbau. 1925, 4, 1129-1135.
Herbert, E. The Measurement of Cutting Temperatures. Proc. of the Instrumentation
of
Mechanical Engineering. 1926, 289-329.
DeVries, M. Drill Temperature as a Drill Performance Criterion. ASTME Technical
Paper
MR68-193. 1968.
Watanabe, K.; Yokoyama, K.; Ichimiya, R. Thermal Analyses of Drilling Process.
Bulletin
of Japan Society Precision Engineering. 1977, 11, 71-77.
Hirao, M. Determining Temperature Distribution on Flank Face of Cutting Tool.
Journal of
Material Shaping Technology. 1989, 6, 143-148.
Agapiou, J.; Stephenson, D. Analytical and Experimental Studies on Drill
Temperatures.
ASME Journal of Engineering and Industry. 1994, 116, 54-60.
Scwerd, F. Ueber di Bestimmung des Temperaturfeldes beim Spanablauf. Z. VDI77.
1933,
211-216.
Boothroyd, G. Photographic Technique for the Determination of Metal Cutting
Temperatures.
British Journal of Applied Physics. 1961, 12, 238-242.
Trent, E. Metal Cutting: Second Edition; Butterworths & Co. Ltd, 1984.
Chao, B.; Trigger, K. Temperature Distribution at the Tool-Chip Interface in Metal
Cutting.
Trans. of ASME. 1955, 1107-1121.
Childs, T.; Maekawa, K.; Maulik, P. Effects of Coolant on Temperature Distribution
in Metal
Machining. Material Science and Technology. 1988, 4, 1006-1019.
Subramani, G.; Kapoor, S.; DeVor, R. A Model for the Prediction of Bore
Cylindricity
during Machining. Journal of Engineering for Industry. Trans. of ASME. 1993, 115
Zheng, Y.; Li, H.; Olson, W.; Sutherland, J. Evaluating Cutting Fluid Effects on
Cylinder
Boring Surface Errors by Inverse Heat Transfer and Finite Element Methods. J. of
Manufacturing Science & Engineering. 2000, 122, 377-383.
Leith, D.; Raynor, P.; Boundy, M.; Cooper, S. Performance of Industrial Equipment
to
Collect Coolant Mist. American Industrial Hygiene Association Journal. 1996, 57
(12),
1142-1148.
Hands, D.; Sheehan, M.; Wong, B.; Lick, H. Comparison of Metalworking Fluid Mist
Exposures from Machining with Different Levels of Machine Enclosure. American
Industrial Hygiene Association Journal. 1996, 57 (12), 1173-1178.
National Institute for Occupational Safety and Health (NIOSH) Health hazard
evaluation
report: Torrington Company, Torrington, Connecticut. 1984.
Gunter, K.; Sutherland, J. An experimental investigation into the effect of process
conditions
on the mass concentration of cutting fluid mist in turning. Journal of Cleaner
Production.
1999, 7, 341-350.
Yue, Y.; Sutherland, J.; Olson, W. Cutting Fluid Mist Formation in Machining Via
Atomization Mechanisms. Proc. of the Symp. on Design for Manufacturing and
Assembly. 1996, 89, 37-46.
Sutherland, J.; Kulur, V.; King, N. An Experimental Investigation of Air Quality in
Wet and
Dry Turning. Annals of the CIRP. 2000, 49 (1), 61-64.
Bayvel, L.; Orzechowski, Z. Liquid Atomization, Combustion: An International
Series;
Taylor & Francis, 1993.
Matsumoto, S.; Belcher, D.; Crosby, E. Rotary Atomizers: Performance Understanding
and
Prediction. Proc. of the 3rd International Conf. on Liquid Atomization and Spray
Systems
at Imperial College at London, 1985.
Weber, C. Zum zerfall eines fl�ssigkeisstrahles. Z Angew Math Mech. 1931, 11, 136-
154.
Matsumoto, S.; Takashima, Y. Atomization Characteristics of Power Law Fluids by
Rotating
Disks. 1st International Conference on Liquid Atomization and Spray Systems. 1978,
145-150.
Kayano, A.; Kamiya, T. Calculation of the Mean Size of the Droplets Purged from the
Rotating Disk. 1st International Conf. on Liquid Atomization and Spray Systems.
1978,
133-137.
Hinds, W. Aerosol Technology; John Wiley & Sons, Inc, 1982; .
Yue, Y.; Sun, J.; Gunter, K.; Michalek, D.; Sutherland, J. Character and Behavior
of Mist
Generated by Application of Cutting Fluid to a Rotating Cylindrical Workpiece, Part
1:
Model Development. Journal of Manufacturing Science & Engineering. 2004, 126, 417-
425.
Sun, J.; Ju, C.; Yue, Y.; Gunter, K.; Michalek, D.; Sutherland, J. Character and
Behavior of
Mist Generated by Application of Cutting Fluid to a Rotating Cylindrical Workpiece,
Part 2: Experimental Validation. Journal of Manufacturing Science & Engineering.
2004,
126, 426-434.
Kinare, S.; Ju, C.; Michalek, D.; Sutherland, J. An Experimental Investigation of
Cutting
Fluid Mist Removal via a Novel Atomizer System. Environmental Sustainability in the
Mobility Industry: Technology and Business Challenges. 2004, SAE SP-1865.
Smoluchowski, M. Versuch einer mathematischen Theorie der Koagulationskinetik
kolloider
L�sungen. Z. Phys. Chem. 1917, 92, 129-168.
Hii, W. W-S. Transient CFD Study of Machining Mist Removal through Kinematic
Coagulation. PhD Dissertation, Michigan Technological University, 2005.
Diniz, A.; Micaroni, R. Cutting Conditions for Finish Turning Process � Aiming: The
Use of
Dry Cutting. International Journal of Machine Tools and Manufacture. 2002, 42, 899-
904.
Weinert, K.; Inasaki, I.; Sutherland, J.; Wakabayashi, T. Dry and Minimum Quantity
Lubrication. Annals of the CIRP. 2004, 53 (2), 511-537.
Dreyer, K.; Hintze, W.; M�ller, M. Schneidstoffe f�r die Trockenbearbeitung.
Werkstatt und
Betrieb. 1997, 130 (6), 420-426.
Byrne, G.; Dornfield, D.; Denkena, B. Advancing Cutting Technology. Annals of the
CIRP.
2003, 52 (2), 483-507.
Porat, R.; Ber, A. New Approach of Cutting Tool Materials � Cermets. Annals of the
CIRP.
1990, 39 (1), 71-76.
Narutaki, N.; Yamane, Y.; Tashima, S.; Kuroki, H. A New Advanced Ceramic for Dry
Machining. Annals of the CIRP. 1997, 16 (1), 43-48.
T�nshoff, H.; Wobker, H.; Brunner, G. CBN Grinding with Small Wheels. Annals of the
CIRP. 1995, 44 (1), 311-316.
Durcros, C; Benevent, V.; Sanchette, F. Characterization and Machining Performance
of
Multilayer of PVD Coatings on Cemented Carbide Cutting Tools. Surface and Coatings
Technology. 2003, 163-164 (30), 681-688.
Cselle, T.; Morstein, M.; Coddet, O.; Geisser, L.; Holubas, P.; Jilek, M.; Sima,
M.; Karimi, A.
LARC: Neue, industrielle Beschichtungstechnologie,. Werkstatt und Betrieb. 2003,
136
(3), 12-17.
Erkens, G.; Cremer, R.; Hamoudi, T.; Bouzakis, K.; Mirisidis, I.; Hadjiyiannis, S.;
Skordaris,
G.; Asimakopouls, A.; Kombogiannis, S.; Anastopoulos, A.; Efstathiou, K.
Supernitrides:
A Novel Generation of PVD Hardcoatings to meet the Requirements of High Demanding
Cutting Applications. Annals of the CIRP. 2003, 52 (1), 65-68.
Cozzens, D.; Olson, W.; Sutherland, J. An Investigation into the Cutting Fluid
Effects on
Temperatures in Aluminum Boring. Proc. of the 1st International Machining and
Grinding
Conference. 1995, 881-891.
Shen, G.; Gandhi, A.; Arici, O.; Sutherland, J. A Model for Workpiece Temperatures
during
Peripheral Milling including the Effect of Cutting Fluids. Transactions of
NAMRI/SME.
2001, 29, 265-272.
Cozzens, D.; Rao, P.; Olson, W.; Sutherland, J.; Panetta, J. An Experimental
Investigation
into the Effect of Cutting Fluid Conditions on the Boring of Aluminum Alloys.
Journal of
Manufacturing Science & Engineering. 1999, 121, 434-439.
Derflinger, V.; Brandle, H.; Zimmermann, H. New Hard/Lubricant Coating for Dry
Machining. Surface and Coatings Technology. 1999, 113 (3), 286-292.
Klocke, F.; Schulz, A.; Gerschwiler, K. Saubere Fertigungstechnolgien � Ein
Wettbewerbsvoteil von morgen? Aachener Werkzeug-maschinen-Kolloquium
Conference Proc., 1996, June 13-14, 4.35-4.108.
Klocke, F.; Gerschwiler, K.; Minimalmengenschmierung � Systeme, Medien,
Einsatzbeispiele und �konomische Aspekte der Trockenbearbeitung, Trockenbearbeitung
vot Matallen. Proc. of the VDI-Seminar. 2003, Stuttgart, Mar. 18: 2.1-2.20.
Karino, K. Developments in Cutting Tools for MQL Cutting. Journal of Japanese
Society for
Tribologists. 2002, 47 (7), 544-549.
Suzuki, S. Developments in Oil Supplying Systems for MQL Cutting. Journal of
Japanese
Society of Tribologists. 2002, 47 (7), 538-543.
Suda, S.; Yokota, H.; Inasaki, I.; Wakabayashi, T. A Synthetic Ester as an Optimal
Cutting
Fluid for Minimal Quantity Lubrication Machining. Annals of the CIRP. 2002, 51 (1),
95-98.
McCabe, J.; Ostaraff, M. Performance Experience with Near-Dry Machining of
Aluminum.
Lubrication Engineering. 2001, 57 (12), 22-27.
N. N., Biodegradable Fluids and Lubricants. Industrial Lubrication and Tribology.
1996, 48
(2), 17-26.
Suda, S.; Yokota, H.; Komatsu, F.; Inasaki, I.; Wakabayashi, T. Evaluation of
Machinability
with MQL System and Effectiveness in Production Lines. Proc. of the International
Tribology Conference. 2001, Nagasaki, Japan, Oct. 29-Nov. 2000, 1, 203-208.
Ju, C.; Keranen, L.; Haapala, K.; Michalek, D.; Sutherland, J. Issues Associated
with MQL
Implementation: Effect on Peripheral Milling Process Performance and Impact on
Machining Economics. IMECE Paper 2005-79259, 2005.
Ackroyd, B.; Compton, W.; Chandrasekhar, S. Reducing the Need for Cutting Fluids in
Drilling by Means of Modulation-Assisted Drilling. Proc. of the ASME. 1998, 405-
511.
Toews, H.; Compton, W.; Chandrasekar, S. A Study of the Influence of Superimposed
Low-
Frequency Modulation on the Drilling Process. Precision Engineering. 1998, (22), 1-
9.
McCabe, J. Advances in Dry Machining of Aluminum. Cutting Tool Engineering.
February
2002.
Filipovic, A.; Sutherland, J., Development of a Magnetostrictive-Actuated Tool
Holder for
Dry Deep Hole Drilling. Transactions of NAMRI/SME. 2005, 33, 437-444.

You might also like