You are on page 1of 7

Experimental Cutting Tool

Mark R. Miller
Los Alamos National Laboratory
Los Alamos, NM 87545
Temperature Distributions
e-mail: mrmiller@lanl.gov
High temperatures in machining cutting zones activate wear mechanisms that decrease
tool life and increase production costs and yet this phenomenon is not fully understood
George Mulholland nor characterized. Although experimental work has been performed, the techniques used
New Mexico State University,
have generally been difficult to apply, and lacked sufficient resolution and or acceptable
Las Cruces, NM 88001
accuracy. Theoretical predictions and computational simulations have been performed to
gain further insight into this problem but could not be accurately validated due to the lack
Charles Anderson of sufficient experimental temperature data. Experimental techniques using modern, digi-
Los Alamos National Laboratory,
tal infrared imaging were developed and successfully applied during this study to gather
Los Alamos, NM 87545
cutting tool temperature distributions from orthogonal machining operations. This new
process has seemingly overcome many problems associated with past experimental
techniques. 关DOI: 10.1115/1.1621425兴

1 Introduction and Background and Fujii to produce temperature distributions between a two-
piece cutting tool. Various metal films were physically vapor de-
Nearly all mechanical work required to remove material during
posited on the inner surfaces between a two-piece cutting tool and
machining is converted to thermal energy that increases the tem- upon observing the melt region after machining tests a single iso-
perature of the work piece, chip and cutting tool. Elevated cutting therm was inferred. This technique requires the use of a variety of
tool temperatures adversely effect material properties leading to films and multiple tests with different tools to produce a single
decreased tool life that correlates to poor work piece geometry temperature distribution. Although this method can be used to
and surface finish. Unfortunately, determination of cutting tool gather temperature distributions for orthogonal machining opera-
temperature distributions is technically difficult and past research tions, the method cannot be used for in situ measurement. Since
has not provided sufficiently accurate temperature data as cited by multiple tests are required to produce a single temperature distri-
Shaw 关1兴, Boothroyd 关2兴, and Kalpakjian 关3兴. Therefore, providing bution, care must be exercised to prevent experimental error from
accurate experimental verification of the mechanisms that cause increasing due to variance between cutting tools and length of cut
wear using modern methods is very important. time.
Industrial demands create problems that require a better under- Kottenstette 关7兴 and Ueda et al. 关8,9兴 used two-color pyrom-
standing of cutting zone temperatures such as increased produc- eters to study cutting zone temperatures. A two-color pyrometer
tion rates to lower costs, dry machining implementation to elimi- 共ratio thermometer兲 uses two detectors sensitive to different wave-
nate hazardous wastes, and continuous product quality lengths to determine target surface temperature from the thermal
improvements. Experimental data regarding cutting tool tempera- radiation. These systems can accurately predict temperature but
ture distributions are required to gain a better understanding of the are restricted to point observation or linear profile, hence an ac-
influence of fundamental machining parameters including cutting curate temperature distribution could not be constructed using this
tool materials and coatings, cutting speed, and work piece mate- ratio thermometer. Ueda 关9兴 also used a two-color pyrometer to
rial. A temperature measurement technique is needed that can pro- predict the maximum temperature of a CBN tool cutting edge at
vide accurate measurements and produce high-resolution thermal the flank face. Ueda correlated the maximum flank face tempera-
images from the cutting zone to study the thermal phenomenon ture to cutting speed and several work piece materials but could
that influences cutting tool wear. Development of a temperature not provide a complete temperature distribution. Experimental
measurement technique that accomplishes these goals was the ob- evidence produced by Boothroyd 关6兴, and Kato and Fujii 关10兴, has
jective of this investigation. shown the maximum cutting tool temperature during orthogonal
A number of techniques have been used to gather temperature machining exists away from the cutting edge along the rake face.
information from the cutting zone of orthogonal machining pro- Stephenson 关11兴 utilized a thermal imaging system to observe
cesses. Early work employed methods such as the cutting tool and average shear plane temperature in the work piece that was used
work piece thermocouple, thermocouple implants, observations of to compare to predictive analytical model results. Mueller-
micro-structural material changes as performed by Trent 关4兴, and Hummel 关12兴 utilized an infrared thermal imaging system to ob-
infrared 共IR兲 photographic plate techniques used by Boothroyd serve the absolute rake face temperature of diamond-coated tools.
关5,6兴. The observations made by Trent to construct temperature Jaspers et al. 关13兴 used a digital infrared camera to study the chip
isotherms required cutting tools constructed of materials that temperatures during machining; unfortunately no cutting tool tem-
would undergo micro-structural changes while experiencing in- peratures were measured. M’Saoubi et al. 关14兴 used a charge
creased temperatures generated during machining, limiting this coupled infrared device to produce cutting zone temperature dis-
technique to a narrow range of cutting tool steels. Boothroyd’s tributions but did not reveal the critical emittance measurement
application of infrared photographic plates was unique but this technique, image resolution nor was an experimental error analy-
early use of infrared technology is now antiquated and no longer sis provided. Additionally, the silicon sensor saturated at tempera-
applicable. Additionally, the photographic plates were sensitive to tures above 900C, therefore critical high temperature information
a narrow range of cutting tool temperature that required preheat- was not obtained.
ing the cutting tool and decreased cutting speeds. Hence, past work does not provide an experimental technique
Physical vapor deposited 共PVD兲 films were employed by Kato to measure accurate and complete high-resolution temperature
distributions for a wide range cutting conditions and cutting tool
Contributed by the Manufacturing Engineering Division for publication in the
materials. A suitable experimental technique utilizing modern
JOURNAL OF MANUFACTURING SCIENCE AND ENGINEERING. Manuscript received digital infrared thermometry was developed during the course of
Dec. 2002; Revised July 2003. Associate Editor: M. A. Elbestawi. this study that overcame many difficulties of measuring cutting

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 667
Copyright © 2003 by ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 2 Work piece showing temperature observation surface

Fig. 1 Orthogonal machining geometry

cutting tool into the end of a rotating cylinder as shown in Fig. 2.


tool temperature distributions during orthogonal machining. The Note the observation area shown in Fig. 2 approximates the target
data is shown to have resolution and accuracy not previously ob- area from which the radiance data was acquired.
tained in past work and the developed technique was applied un- Radiometers measure radiant flux, or radiance, from a target
der ordinary shop floor conditions. The digital infrared thermom- surface and generate a corresponding output signal. This electric
eter used is sensitive to thermal radiation corresponding to signal, usually a voltage, is converted to temperature using a cali-
temperatures between 0 and 2000 C allowing observation of a bration algorithm that is derived under ideal conditions from a
wide range of cutting speeds. Since nearly all cutting tool materi- blackbody surface. Generally at this point an emittance value,
als and associated coatings in use today emit thermal radiation input by the user, is applied to the data to correct for ‘‘real’’
proportional to temperature the application of IR thermometry as surfaces 共emittance less than one兲. Once these calculations are
described in this study offers wide application. This method also complete a thermogram is generated that maps the apparent tem-
allows real time data acquisition that can be readily correlated to perature of the target area. The total radiant flux as seen by a
other data obtained from instrumentation, such as force transduc- detector is comprised of target surface emission and reflectance
ers, observed over the same period. from surroundings as well as atmospheric effects such as absorp-
Experimental temperature data is needed to validate and im- tion, emission, and scattering.
prove computational models capable of predicting tool tempera- As stated all radiometers require knowledge of target surface
ture and the mechanisms that cause tool wear. The models in emittance to compensate for real surfaces that radiate less effi-
conjunction with experimental data may be used to help select ciently than black body surfaces. Additionally, the thermal radia-
cutting tool materials and coatings, design tool geometry and op- tion emitted by most real surfaces is dependent on surface tem-
timize machining processes. This benefits industry by extending perature as well as other factors including material composition
cutting tool life that reduces production costs, potentially increas- and surface texture. Hence, accurate emittance values that reflect
ing production rates and improving part quality. Cutting zone test conditions of the target surface are directly related to the
simulation, provided by finite element analysis, yields complete accuracy of the inferred temperature results. Since published data
temperature distributions; unfortunately past studies have not de- may not reflect actual test conditions of the target surface 共tem-
veloped methods or provided data to make these comparisons. perature, roughness, oxidation, etc.兲 using it to infer true tempera-
Modern digital infrared thermal imaging developed during the ture from radiation thermometers may not yield accurate results.
course of this study has demonstrated the necessary accuracy and Target surface emittance measurements are critical when accurate
resolution to allow valid comparisons between experimental and surface temperatures are required using radiation thermometers
computational results. and must be performed using target surface conditions 共tempera-
Although orthogonal machining principles have been exten- ture, finish, etc.兲 that are as close as possible to those existing
sively used for simulation and testing in past work, a thorough during actual testing.
understanding of this process has not been realized. The orthogo- Madding 关15兴 illustrated the mathematics behind radiation ther-
nal machining technique offers many benefits for testing including mography that was employed during this study. Radiance as mea-
a reduction of experimental variables, profuse existence of theory sured by an IR camera, also termed camera digital level, L m , is
and models, and offers valuable insight to oblique machining. given by Eq. 共1兲
Orthogonal machining tests can potentially be used to examine L m ⫽␧ t L t ⫹ 共 1⫺␧ t 兲 L b (1)
how cutting tool materials and coatings as well as rake angle and
edge preparation, effect cutting zone temperatures. The emittance of the target surface is represented by ␧ t , and
hence ␧ t L t is target surface radiance and (1⫺␧ t )L b is the radiance
reflected off the target from background sources. The radiance
2 Theory measurements produced by the infrared camera used in this study
Infrared imaging requires an unobstructed view to the target are based on the assumption that the emittance of the target sur-
surface of interest, hence orthogonal machining was employed face is independent of wavelength, i.e., a greybody emitter as
during all tests. Figure 1 depicts simple orthogonal machining verified by Madding 关16兴. Therefore, 1⫺␧ t ⫽ ␳ t , is used to de-
geometry where the cutting tool has a rake angle represented by a scribe the reflectivity of the target surface. Prior to testing back-
and the primary shear plane angle is designated by ␾. The depth ground temperatures were measured with the radiometer and
of cut is t 0 and the chip thickness after deformation is t 1 . found consistent with normal room conditions 共approximately 24
Orthogonal machining is produced by holding the cutting edge C兲. Note Eq. 共1兲 does not include potential atmospheric effects.
of a tool perpendicular to the direction of relative motion between Equation 共1兲 can be solved to yield an expression describing target
the cutting tool and work piece. This machining geometry pro- level radiance, Eq. 共2兲.
duces a continuous, rectangular chip and generates a plane surface
L m ⫺ 共 1⫺␧ t 兲 L b
parallel to the original work surface. Orthogonal machining, as L t⫽ (2)
illustrated in Fig. 1, was demonstrated for this study by feeding a ␧t

668 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 3 Cutting zone experimental test setup Fig. 4 Emittance data

The radiometer employed in this study used Eq. 共2兲 to calculate


the digital radiance value for the target surface and determine the The target surface of interest is the vertical side of the cutting
temperature from a calibration 共lookup兲 table based on data gath- tool directly below the rake surface. Two regions generate heat in
ered under ideal conditions. the cutting zone of an orthogonal machining operation, the pri-
mary and secondary zone and are depicted in Fig. 1. Plastic de-
formation causes almost instantaneous heating of the material in
3 Experiment the primary zone while both viscous heating and friction contrib-
The experimental setup shown in Fig. 3 reveals the location and ute to heat generated in the secondary zone. The mechanisms of
orientation of the infrared camera in relation to the cutting tool heat transfer include conduction, convection and thermal radiation
and work piece. The mild steel work piece, AISI 1025, has pub- absorption and reflectance. The techniques described in this work
lished material properties as follows, elastic modulus equal to are based on several fundamental assumptions from past work and
203.37 GPa, Poisson’s ratio of 0.303 and hardness approximately observed phenomenon during testing. First, the target surface acts
140 HB. The cutting tool is fine grain carbide, Kennametal, Inc. as a greybody emitter as verified by Madding 关16兴. Next, Boo-
grade K313. Grade K313 is advertised as an unalloyed WC/Co throyd 关2兴 found convection losses from the target surface are
with low binder content for application as an ISO class M10 to negligible during steady state machining conditions and are omit-
M20 or K05 to K20 but the composition is most similar to K20. ted throughout this study. Also, three pathways exist for heat con-
Hardness is approximately 1900 HV with an elastic modulus duction in the cutting zone, a兲 into the work piece, b兲 into the
equal to 615 GPa and a Poisson’s ratio of 0.216. The cutting tool chip, and c兲 into the cutting tool. Lastly, the amount of heat con-
incorporated a 15° positive rake angle and a chip deflector to help ducted into the work piece in the direction of motion is insignifi-
move the chip away from the camera lens. This geometry pro- cant at the practical cutting speeds as used for these tests as origi-
duced a continuous, rectangular chip for all tests conducted. nally discovered by Rapier 关18兴.
The selection of cutting speeds and depth of cut was intended to The IR camera was set on broad band spectrum, 3–14 mi-
represent general machining conditions undergoing final material crometers, that brackets the peak spectral radiance for the tem-
removal conditions or finish cuts. Table 1 yields the test param- perature detected during testing and an emittance of one was used
eters for the temperature distributions gathered. The cutting as IR camera input to reduce subsequent calculations. The camera
speeds are based on recommended values found in the Machin- was setup normal to the target surface as shown in Fig. 3 due to
ery’s Handbook 关17兴 for turning a mild-steel work piece with a the directional nature of the thermal radiance from the target
standard carbide cutting tool. The 0.102 mm 共0.004 inch兲 per surface.
revolution feed rate correlates to a 0.102 共0.004 inch兲 depth of cut The reference temperature method was used to develop an ex-
for orthogonal machining. pression for target surface emittance as a function of camera level
Plane strain conditions exist in the chip during orthogonal ma- radiance. An example of emittance testing is shown in Fig. 4 and
chining in the same plane created by the cutting and thrust forces. represents an average of six data points from ten cutting tools
The degree of strain in the lateral direction, perpendicular to this each. The figure also shows the linear regression line along with
plane, can be determined from the ratio of the work piece wall its associated 95% confidence interval. Multivariate cubic spline
thickness to the chip width after material is removed. The lateral interpolation was used to produce the constants shown in Eq. 共3兲;
strain for this case was less than 0.5% as tested. this equation was used to plot the linear curve shown in Fig. 4.

Table 1 Mechanical data and test parameters

RPM rev/min 550 760 1050


Feed Rate mm/rev 共inch/rev兲 0.102共0.004兲 0.102共0.004兲 0.102共0.004兲
Depth of Cut mm 共inch兲 0.102共0.004兲 0.102共0.004兲 0.102共0.004兲
Cutting Speed m/s 共ft/min兲 1.78 共351兲 2.46 共485兲 3.40 共671兲
Chip Thickness mm 共inch兲 0.203共0.008兲 0.203共0.008兲 0.178共0.007兲
Chip Width mm 共inch兲 4.67共0.184兲 4.67共0.184兲 4.70共0.185兲
Strain % 0.5 0.5 1.1
Chip Contact Length mm 共inch兲 0.635共0.025兲 1.092共0.043兲 0.762共0.030兲

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 669

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 5 Test results observed at a cutting speed of 1.78 mÕs Fig. 7 Test results observed at a cutting speed of 3.40 mÕs
„351 ftÕmin…, plotted as isotherms showing cutting tool outline „671 ftÕmin…, plotted as isotherms showing cutting tool outline

Observation of Fig. 4 reveals the dependence of emittance on


target surface radiance. Considering that temperature is propor-
tional to radiance, Fig. 4 illustrates obtaining accurate emittance ture values. A random sample of images was chosen from each
measurement is essential to the accurate formulation of tempera- segment of tape between 1.0 and 5.4 seconds after test start and
ture distributions. averaged to create a matrix of temperature values. Each numeric
element of the averaged data matrix corresponds to a single pixel
␧ t 共 L m 兲 ⫽0.00001793L m ⫹0.192 (3) within the image.
␧ t (L m ) in Eq. 共3兲 represents the target surface emittance as a As stated, the radiometer used in this study provides a tempera-
function of camera digital level. Equation 共3兲 is placed in Eq. 共2兲 ture value for each matrix element of a thermogram. Utilization of
to yield the following. Eq. 共4兲 requires camera level radiance to predict the desired result,
target surface radiance. The following procedure was used effec-
L m ⫺ 共 1⫺␧ t 共 L m 兲兲 L b tively to apply Eq. 共4兲 to the average data matrix. First, convert
L t⫽ (4) each element of the averaged thermogram matrix from tempera-
␧ t共 L m 兲
ture to camera level radiance. Next, apply Eq. 共4兲 to each element
Equations 共3兲 and 共4兲 were applied to the target surface radiance to correct for emittance as a function of temperature to obtain true
data used to produce the image of the target surface shown in target radiance values. Finally, convert target radiance to target
Figs. 5, 6, and 7. temperature for the final images shown in Figs. 5, 6, and 7. The
Radiance data gathered for these machining tests were recorded temperature to radiance and radiance to temperature conversions
on VHS videotape at the rate of 30 Hz for the period of testing, were performed using second order polynomials fit to the radiom-
approximately six seconds. Each frame of data comprised a com- eter manufacturer supplied camera calibration data. A commer-
plete black and white image of the target surface area as observed cially available calculation software package was used to develop
by the radiometer and is stored digitally as a matrix of tempera- the conversion equations and apply Eq. 共4兲 to the data. Conversion
accuracy required the use of second order polynomials best fit to
neighborhoods of the calibration data, hence the coefficients are
too numerous to list. The numerical error of converting from ra-
diance to temperature and back to temperature was found to be
negligible.
Prior to testing, each cutting tool was prepared and inspected.
Preparation included grinding, lapping the target surface, and
placement of two fiducial marks, 0.254 mm 共0.010 inch兲 diameter
by 0.254 mm 共0.010 inch兲 deep, by electronic discharge machin-
ing on the target surface. During initial cutting tests, observation
of the target surface revealed the carbide cutting tool was becom-
ing darker at the cutting edge. Since, accuracy could be affected
by a change in surface properties during testing the cutting tools
were preheated in a furnace to over 500°C to uniformly darken the
target surface and prevent any emittance change during the test.
Isotherm images do not, in general, represent spatial detail well;
consequently a method was derived and employed to reproduce
the cutting tool outline of the image. Each cutting tool was in-
spected before testing using a video microscope capable of accu-
rate positioning to within 0.0025 mm 共0.0001 inch兲 and this in-
formation was used to reconstruct the outline of each tool using
CAD software. The final isotherm was then imported into the
Fig. 6 Test results observed at a cutting speed of 2.46 mÕs CAD file where the cutting tool outline was located. The isotherm
„485 ftÕmin…, plotted as isotherms showing cutting tool outline image was then located and scaled using the two fiducial marks

670 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 8 Target temperature error as a function of emittance
error Fig. 9 Target temperature as a function of target temperature
error

that are readily observed in the thermal image. Additionally, the


contact length of the chip was measured using the video micro- Application of the root-sum square method was also used to
scope after each test. find target temperature uncertainty. Equation 共7兲 is used to ap-
proximate a radiantly linear system.
T⫽A ln共 L t 兲 ⫹B (7)
4 Results and Data Analysis
Equation 共8兲 describes the temperature error as a function of target
The reduced experimental temperature data, gathered from the radiance level.
infrared images of the target surface, were plotted as isotherms as
shown in Figs. 5, 6, and 7 with contour lines separated in 25 C A⌬L t
increments. The cutting tool outline and the work piece are clearly ⌬T⫽ (8)
Lt
visible, depicted in red, with the cutting tool edge appearing in the
upper right corner. The work piece is to the right of the vertical Target level uncertainty was found by application of the root-sum-
lines. The uncut surface is represented by the vertical line extend- square method to the radiosity Eq. 共2兲 and using Eq. 共8兲 the total
ing to the top of the figures while the vertical line extending from expected target temperature uncertainty was found to be described
the cutting tool edge to the bottom of the figures represents the by Eq. 共9兲, where the constant A is derived from Eq. 共7兲.
newly cut surface. The rectangular shape located just above the
cutting tool edge on the rake face represents the chip thickness
and contact length.
⌬T t ⫽A
冑 ⌬L m
2
⫹ 冉 冊
⌬␧ t
␧t
2
共 L m ⫺L b 兲 2 ⫹⌬L 2b 共 ␧ t ⫺1 兲 2
(9)
An estimate of resolution and scale is derived from the small ␧ tL t
red circles, representing the cutting tool fiducial marks, located in
Equation 共9兲 illustrates total target temperature measurement un-
the lower left corner of the isotherm images. As measured from
certainty due to emittance, camera level and background level
the image the pixel height and width is 0.018 mm 共0.0007 inch兲
errors. Equation 共9兲 shows that as emittance decreases, tempera-
by 0.015 mm 共0.0006兲 inch respectively.
ture error increases, and as target level radiance increases, total
An error budget was performed on both the emittance measure-
target temperature error decreases. Using Eq. 共9兲 and the same
ments and the cutting tool target temperature results following
radiance values as before, a plot of total target temperature error
Madding’s 关15兴 methodology. The reference temperature method
as a function of target temperature was plotted and is shown in
used to gather emittance data as described by the following
Fig. 9.
equation.
L t ⫺L b 5 Discussion
␧ t共 L m 兲 ⫽ (5)
L m ⫺L b Test results are shown for three cutting speeds, 1.78 共351兲, 2.46
Where L t is target surface radiance, L m is measured camera level 共485兲, and 3.40 m/s 共671 ft/min兲, in Figs. 5, 6, and 7 respectively.
and L b is reflected background radiance. Equation 共6兲 represents The isotherm plots provide an accurate, well-defined temperature
emittance fractional error and was obtained by application of the distribution over the cutting tool target surface. Accuracy of the
root-sum-square method to Eq. 共5兲. results from Fig. 5 is illustrated in Fig. 9, where as temperatures

冑 冋冉 冊冉 冊册
reach 500 C the total temperature error is approximately 5 C. This
⌬␧ t 1 1 figure reveals an important and interesting trend; as temperature
⫽ 2⌬L 2 ⫹ (6) increases, target temperature error decreases. Intrinsically, all ra-
␧t 共 L t ⫺L b 兲 2
共 L m ⫺L b 兲 2
diation thermometers produce this phenomenon. Figure 10 de-
Emittance fractional error was derived from Eq. 共6兲 using radi- scribes target temperature as a function of emittance measurement
ance data from cutting tool target surface shown in Figs. 5, 6, and error and shows that at a temperature of 200 C the error is ap-
7. The manufacturers quoted a calibration specification of ⫾0.1°C proximately 4%, dropping to below 1% as the target temperature
for the camera used in this study. Therefore, ⌬L, was determined increases. Examination of Fig. 10 shows emittance measurement
from radiance value calculated for 0.2°C multiplied by 3.0 to ap- error had a diminishing effect on target temperature error. Based
proximate a three sigma standard deviation for the temperature on these observations, the radiation thermometer and test proce-
range of the experiments. The emittance measurement error was dure used in this study provided accurate temperature data. Since
obtained using Eq. 共6兲 and radiance levels extracted from the cut- these trends are inherent to radiation thermometers, they make an
ting tool target surface shown in Fig. 5; These errors are plotted as attractive instrument to use for measurement of high temperature
a function of target temperature in Fig. 8. regions as encountered in machining cutting zones.

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 671

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


nitride have produced emittance values from 0.8 to 0.9. Although
these emittance values will undoubtedly differ for other cutting
tools constructed of similar materials they demonstrate the wide
variety of cutting tool materials that may be accurately observed
using digital infrared thermal imaging.

6 Conclusion
Digital infrared thermal imaging applied to orthogonal machin-
ing with the techniques developed during the course of this study
can be used to gather temperature distributions for a wide variety
of cutting tool materials and cutting speeds. The error analysis
performed on the temperature data gathered during this study pro-
vides evidence the isotherms accurately depict the temperature
distributions. This evidence also demonstrates the application of
digital infrared imaging can solve many of the problems associ-
ated with past experimental techniques. The high-resolution
共0.018 mm by 0.015 mm兲 temperature distributions were shown to
Fig. 10 Target temperature as a function of emittance error be accurate within 1% at higher temperature values and therefore
provide for precise validation of computational simulations. Ad-
ditionally, this technique should be applicable to almost any cut-
ting tool material with measurable emittance. Also digital infrared
The temperature resolution produced during testing as depicted
imaging was shown to become more accurate as temperature in-
in Figs. 5, 6, and 7 allows for meaningful comparison with past
creases, hence accurate temperature distributions obtained from
orthogonal metal cutting simulation. Strenkowski and Moon 关19兴
orthogonal machining at higher cutting speeds, producing gener-
applied the finite element method to construct an Eulerian model
ally increased temperatures, may be captured and studied. Digital
of an orthogonal machining operation. The smallest element of the
infrared thermometry has limitations however. The method must
model was approximately 0.0254 mm by 0.1778 mm 共0.001 inch
have an unobstructed line of sight measurement from the camera
by 0.007 inch兲 as observed from their results. The resolution pro-
to the target surface and therefore generally not suitable for ob-
duced by the infrared imaging for this study is approximately
lique machining applications. Obtaining critical emittance mea-
0.0152 mm by 0.0178 mm 共0.0006 inch by 0.0007 inch兲 and well
surements is involved and generally requires target surface prepa-
within this range. The data produced in this study precludes any
ration. Additionally infrared thermometry can only be applied in
temperature comparison with Strenkowski and Moon due to the
dry machining applications as the coolant obstructs the view of
different machining conditions and materials. However, new mod-
the cutting tool target surface.
els could be constructed or validation tests performed that would
Nevertheless infrared digital thermometry as applied to or-
allow direct comparison.
thogonal dry machining offers the potential to study cutting zone
The spatial detail of the carbide cutting tool was accurately
temperature distributions and hence may be used optimize many
reproduced within the thermal images represented in Figs. 5, 6,
critical cutting tool parameters such as rake angle, edge prepara-
and 7. These figures also reveal a cavity effect that is a conse-
tion, surface roughness, and coatings. High temperatures as en-
quence of digital infrared thermal imaging. Two cavities exist in
countered in the cutting zone of machining processes affect mate-
the images, 1兲 between the clearance face and work piece, and 2兲
rial properties and in most cases reduce tool life. Since cutting
along the rake face where the chip looses contact with the cutting
tool life directly impacts machining efficiency and production
tool. These cavities produce multiple reflections of radiant energy
costs understanding the effects of cutting tool geometry, materials,
from within an enclosure that results in an apparent increase in
coatings, and edge preparation is very important. Digital infrared
emittance as described by Kaplan 关20兴. Accordingly, the isotherms
thermal imaging as applied in this study can be used to gather
could be extrapolated to these surfaces with confidence. Since no
accurate, high-resolution temperature distributions from orthogo-
cavity exists in the most critical area, along the rake face in con-
nal dry machining.
tact with the chip, these temperatures are accurately revealed.
Observations made from the isotherm temperature distributions
shown in Figs. 5, 6, and 7 reveal as expected that generally in- References
creased temperatures are produced as cutting speed increases. Fig- 关1兴 Shaw, M. C., 1984, Metal Cutting Principles, Oxford University Press.
ure 5 represents a cutting tool temperature distribution obtained at 关2兴 Boothroyd, G., 1989, Fundamentals of Machining and Machine Tools, 2nd ed.,
Marcel Dekker.
a cutting speed of 1.78 m/s 共351 ft/min兲 with 550 C the highest 关3兴 Kalpakjian, S., 1991, Manufacturing Processes for Engineering Materials, 2nd
recorded isotherm. Figure 6 shows 675 C as the highest recorded ed., Addison-Wesley.
isotherm produced at a cutting speed of 2.46 m/s 共485 ft/min兲 and 关4兴 Trent, E. M., 1977, Metal Cutting, Butterworths.
Fig. 7 shows 775 C the highest recorded isotherm at 3.40 m/s 共671 关5兴 Boothroyd, G., 1961, ‘‘Photographic Technique for the Determination of Metal
Cutting Temperatures,’’ Br. J. Appl. Phys., 12, p. 238.
ft/min兲. Each of the figures show the highest temperatures located 关6兴 Boothroyd, G., 1963, ‘‘Temperatures in Orthogonal Metal Cutting,’’ Proc.
away from the rake face and the contact area between the cutting IME, Vol. 177, p. 789.
tool and chip. One explanation for this phenomenon is that the 关7兴 Kottenstette, J. P., 1986, ‘‘Measuring Tool-Chip Interface Temperatures,’’
chip conducts heat from the area of contact with the cutting tool ASME J. Eng. Ind., 108, May, pp. 101–104.
关8兴 Ueda, T., Sato, M., and Nakayama, K., 1998, ‘‘The Temperature of a Single
and this heat is carried away as the chip passes over the tool. This Crystal Diamond Tool in Turning,’’ CIRP Ann., 47, pp. 41– 44.
would produce higher temperature regions within the cutting tool 关9兴 Ueda, T., Huda, M. A., Yamada, K., and Nakayama, K., 1999, ‘‘Temperature
away from the rake face. Measurement of CBN Tool in Turning of High Hardness Steel,’’ CIRP Ann.,
As shown, this study provided accurate temperature data for an 48, pp. 63– 66.
关10兴 Kato, T., and Fujii, H., 1996, ‘‘PVD Film Method for Measuring the Tempera-
orthogonal machining operation performed at recommended cut- ture Distribution in Cutting Tools,’’ ASME J. Eng. Ind., 118, February, pp.
ting speeds on a steel work piece with a carbide cutting tool. The 117–122.
equipment and procedures used to perform these tests should be 关11兴 Stephenson, D. A., 1991, ‘‘Assesment of Steady-State Metal Cutting Tempera-
applicable to a wide variety of work piece and cutting tool mate- ture Models Based on Simultaneous Infrared and Thermocouple Data,’’ ASME
J. Eng. Ind., 113, May, pp. 121–128.
rials at almost any cutting speed. The emittance of TiN coated 关12兴 Mueller-Hummel, P., and Lahres, M., 1995, ‘‘Temperature Measurement on
cutting tools have been measured and found to be in an acceptable Diamond-Coated Tools During Machining,’’ Industrial Diamond Review, 295,
range from 0.1 to 0.2. Other cutting tool materials such as boron February, pp. 78 – 83.

672 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


关13兴 Jaspers, S. P. F. C., Dautzenberg, J. H., and Taminiau, D. A., 1998, ‘‘Tempera- 关18兴 Rapier, A. C., 1955, ‘‘A Theoretical Investigation of the Temperature Distri-
ture Measurement in Orthogonal Metal Cutting,’’ Int J Adv Manuf Technol, 14, bution in the Metal Cutting Process,’’ Br. J. Appl. Phys., 5, pp. 400– 407.
pp. 7–12. 关19兴 Strenkowski, J. S., and Moon, K. J., 1990, ‘‘Finite Element Prediction of Chip
关14兴 M’Saoubi, R., Lebrun, J. L., and Changeux, B., 1998, ‘‘A New Method for Geometry and Tool/Workpiece Temperature Distributions in Orthogonal Metal
Cutting Tool Temperature Measurement Using CCD Infrared Technique: In- Cutting,’’ ASME J. Eng. Ind. 112, November, pp. 313–318.
fluence of Tool and Coating,’’ Mach. Sci. Technol., 2, pp. 369–382. 关20兴 Kaplan, H., 1993, Practical Applications of Infrared Thermal Sensing and
关15兴 Madding, R. P., 1999, ‘‘Emissivity Measurement and Temperature Correction Imaging Equipment, SPIE Optical Engineering Press, Vol. TT 13.
Accuracy Considerations,’’ Proc. SPIE, 3700, pp. 393– 401. 关21兴 Madding, R. P., Guyer, E. C., and Brownell, D. L., 1989, Handbook of Applied
关16兴 Madding, R. P., 1982, ‘‘Science Behind Thermography,’’ Proc. SPIE, 371, pp. Thermal Design; Infrared Thermography, McGraw-Hill.
2–9. 关22兴 Miller, M. R., 1999, ‘‘An Experimental Study of Cutting Tool Temperature
关17兴 Greene, R. E., Editor, 1996, Machinery’s Handbook, Industrial Press Inc., 25th Distributions Generated During Orthogonal Machining,’’ Ph.D. thesis, New
ed. Mexico State University, Las Cruces, NM, 1999.

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 673

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 01/28/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like