You are on page 1of 10

Using High-Speed Infrared Imaging as a Tool to Design Materials with

Further Improved Machinability


I. Arriola Aldamiz(1)(2), P.J. Arrazola(1), M.A. Davies(3), A. Cooke(3) and B. Duttered(3)
(1)
Mondragon University
(2)
CicMargune
(3)
University of North Carolina at Charlotte

ABSTRACT
Materials’ cutting during machining is an extremely complex thermo-mechanical problem due to the severe
physical conditions associated to materials shearing mechanisms. During chip formation very high elasto-
viscoplastic strains of the order of 1-2, strain rates higher than 104 s-1; temperature increase from room
temperature to 1000 ºC, and temperature rates up to the order of 106 ºCs-1 occur in milliseconds.
Furthermore, all these phenomena occur in an area smaller than 1mm x 1mm. In this work, we have first tried
to measure the micro-scale temperature fields on tool-chip temperature accurately during the cutting of two
AISI 4140 steels with different machinability ratings and secondly, tried to distinguish the two steels.
Orthogonal cutting experiments were performed on a high speed machining center with surface speeds up to
400 m·min-1 and uncut chip thicknesses of 0.2 mm. To accomplish this, a high-bandwidth infrared imaging
system from Indigo Systems, Inc., that utilizes a 320 by 256 Indium-Antimonide liquid-nitrogen-cooled,
detector array was used. The tests were conducted at 488 frames per second, variable exposure time for
each condition and a resolution of the system of 5-10 micrometers. The results indicate that in the rake face
as well as flank face of the thermal field, improved machinability correlates with significant reductions in
temperature that exceed measurement uncertainties of approximately 45ºC. Such micro-scale temperature
measurements will help to design materials with further improved machinability.

Keywords: High-speed thermal imaging, micro-escale temperature, variable integration time, machinability

INTRODUCTION
Modern manufacturing industry is making increasingly severe demands on the technological and operating
characteristics of steels because the machining costs, in industrialised countries such as USA, are more than
15% of the cost of all the goods produced by all the manufacturing industry [1]. To satisfy these demands it is
necessary to: 1) improve steels’ properties to reduce the consumption of metal and increase the sustainability
of machined parts, and 2) extend the use of automatic production lines, increasing labor productivity, and
reducing the cost of cutting tools, labor, coolants and energy costs [2].

This cost reduction can be achieved by either a) developing new tool types; shape and coatings [3], b) using
minimum quantity lubrication coolants (MQL) [4] or c) developing new type of materials by means of different
treatments or adding different elements to enhance their machinability rate [5].

In an effort to understand the machining process in more detail, many researchers have tried to measure
cutting forces, surface roughness, chip dimensions, strain, strain rates, tool wear and temperature during
orthogonal cutting. Among these, temperature is claimed to be the most difficult to measure, which explains
the number of different methods used over the years [6].

There are three main sources of heat generation during the process of cutting metal with a machine tool: I)
(source 1) Irreversible plastic deformation. Under the shearing force from the tool, the chemical and
physiochemical bonds along the shearing region are broken, II) (source 2) Friction and shear on the tool
rake face, or secondary shear zone and III) (source 3) Friction created between the flank face of the tool and
the new surface of the work material. Heat from these three sources is converted from the 95-99% cutting
energy which is used to cut the workpiece. As a result of this heat, the tool, chip, and work material elevate in
temperature.

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


1 source chip

2 source
3 source

workpiece tool
Primary shear zone Secondary shear zone
εeq=1-2 εeq=2- 4
εeq=104-105 s-1 εeq=104-105 s-1
T= 400-600 K T= 800-1000 K

Figure 1. Orthogonal cutting: assumes that the cutting edge of the tool is set in a position that is perpendicular
to the direction of relative work or tool motion. This allows us to deal with forces that act only in one plane.

A lot of different techniques have been employed to measure these machining temperatures. For example,
calorific methods [7], Fibre Bragg gratings [8], Single wire thermocouples [9], resistance methods [10],
thermal sensitive paints [11], tool-work thermocouple [12], embedded thermocouple [13], pyrometer [14] and
infrared cameras [15] are among the most important or most used ones. Each of these methods has their
advantages and disadvantages, but nowadays the infrared technique is at its peak. This non-invasive method
provides high frame rate, variable integration times, considerably small spatial resolution, macro/micro
analysis, temperature analysis through time, etc.

Nevertheless, the main disadvantage of this infrared methodology during machining is that the temperatures
reached in the center of the chip are not possible to measure. It has been observed that the temperature
gradient is higher across the chip cross section than along the length of the chip. Thereby, it is believed that
the maximum temperature during machining is reached in the center of the contact length between the chip
and tool. So, it has to be kept in mind that in these experiments, the tool and the resulting chip will be
measured from the side. However, it will be possible to analyze the tool-chip-workpiece system in detail and
obtain valuable temperature data. Later, using inverse simulation, the maximum temperature in the cross-
section of the chip could be assessed. This is complex, cumbersome and time-consuming due to many
parameters involved: material law behavior, friction coefficient, strains, strain rates, heat fluxes, etc. In other
words, many variables and assumptions playing critical roles in the Finite Element Method (FEM). This would
lead to an increase in the uncertainty of the measurements. Therefore, the temperatures obtained in the side
are reliable enough to understand the machining process, so researchers work with them.

Figure 2 shows an image taken with a high speed camera and a thermal map with an infrared camera as an
illustration of what is going to be measured from the side.

AIR
WORKPIECE
WORKPIECE AIR

CHIP
CHIP

TOOL TOOL
100µm

Figure 2. High-speed imaging in the visible spectrum and thermal camera picture in the infrared spectrum.

The objective of this paper is to measure the temperatures reached during the machining of two AISI 4140
steels with different machinability rates under different cutting conditions. First, it will be attempted to obtain
temperatures as accurate as possible and secondly, try to distinguish the two materials thermally.

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


EXPERIMENTAL SET UP
The thermal imaging microscope was a high-bandwidth infrared imaging system from Indigo Systems, Inc.
that utilizes a 320 by 256 pixels, liquid-nitrogen-cooled, Indium-Antimonide detector array and accepted mid-
range radiation from 3 µm to 5 µm in wavelength. The experiments were accomplished with the maximum
frame size in order to obtain the whole workpiece-tool-chip system, so limiting the bandwidth to 488 frames
per second. The resolution of the system was nearly wavelength limited and is approximately 5-10
micrometers. The IR camera, the microscope (15x), an aluminum monolithic tool post, a 12.5 mm square tool
holder, and a zero-rake, five-degree clearance, tungsten carbide insert were attached to a 37 mm thick
aluminum plate. At the same time, the plate was attached to the Makino A55 high-speed machining centre to
produce an orthogonal cut by removing the end of the tube.

Tool insert
Monolothic Monolothic
Workpiece
toolpost toolpost
20,000
Microsope rpm
body spindle
Machine
Z-axis
Tool insert

Microsope
body
Infrared
camera
Monolothic
toolpost

Figure 3. Makino A55 high-speed machining centre (left) and experimental set up of the IR camera (right).

Each workpiece was a tube 45-mm long, 25-mm diameter, 1-mm thick, tubular section, and a 30 mm long
sold base. The tool was custom ground so that the desired image plane of the tool was perpendicular to the
optical axis of the microscopy system. Orthogonal cutting experiments were performed for a wide range of
cutting parameters while varying the cutting speeds (v) up to 400 m·min-1 and the uncut chip thickness (h) of
0.2mm.

CALIBRATION SET UP
In this work, both workpieces as well as the tool insert were calibrated in order to obtain the temperature of
the whole workpiece-tool-chip system. Both samples were modified to fit a cylindrical electric heater to
produce a uniform thermal contact. On one hand, the workpiece material was machined to obtain an inner
diameter of 12.7 mm and a wall thickness of 1 mm and on the other hand, the tool insert was EDM a hole in
the center. During the calibrations, the microscope was first focused on the artifact surface and then, set
voltages were applied to the heater and held for several minutes to stabilize the samples’ temperature. For
each integration time use in the experiments, the temperature was varied from 50ºC to 750ºC in intervals of
50ºC. Each temperature was controlled with a calibrated type-K thermocouple and simultaneously the
camera’s signal (S) was recorded.

Figure 4. Calibration set up for the workpieces (left) and tool insert (right).

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


The signal of the camera versus temperature from the thermocouple was fitted using an interpolating function
shown below [15].

T ≡ α1. ln(S + α 2 ) + α 3

where α1, α 2 and α 3 are parameters that were varied to produce the least squared curve-fit for each data set.

Map each pixel using


integration time

Figure 4. Calibration curve plot (bottom)

Two things have to be pointed out regarding these calibrations that later, during the temperature and
uncertainty assessment, have to be taken into account. First, thermal imagers have non-uniform spatial
responsivity as well as the non-linear responsivity of the individual detecting elements. Therefore, one unique
approach taken in this work to minimize errors was to disallow the use for a non uniformity correction (NUC)
in the camera software and calibrate each pixel as an independent measuring device. Thus, the interpolating
functions act both to calibrate each independent pixel signal in terms of radiance temperature, and also to
remove the image non uniformity resulting from detector variation. Secondly, since we had no black body, the
samples were peroxidised before conducting the calibration. The aim of this was to obtain calibration
conditions close to the black body. In fact, similar conditions were obtained as it can be seen in Figure 6; a
peroxidised body has a high emissivity value of at least 0.9.

TEMPERATURE AND UNCERTAINTY ASSESSMENT


Depending on the emissivity uncertainty and the type of the sensor of the camera that is being used during
the measurements or calibrations, there can be considerable uncertainties in temperature as can be seen in
Figure 5.

a) b)
Figure 5. (a) Uncertainty in temperature due to different spectral bands [16] and (b) uncertainty in temperature for a
surface of 0.3 of emissivity at 1000 Celsius using the Indium-Antimonide detector (3 µm to 5 µm).

Therefore, in order to reduce as much as possible the effect of the emissivity, all the samples that were used
in the experiments were introduced in a vacuum Fourier transform infrared (FT-IR) spectrometer. Thereby,

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


accurate emissivity data for each sample were obtained as it is plot in Figure 6. This graph shows how the
temperature, wavelength, surface roughness, oxidation and the material can affect on the emissivity.
Consequently, the temperature uncertainty can be considerable.
1
Tool_oxi_502ºC
0.9
Tool_oxi_699ºC
0.8 4140_oxi_301ºC
0.7 4140_oxi_697ºC
Emissivity (ε)

0.6 Samples with oxidation

0.5 Samples polished


0.4 Tool_306ºC
Tool_699ºC
0.3
Tool_Ra:63nm_68ºC
0.2 Tool_Ra:150nm_689ºC
0.1 4140_696ºC
4140_299ºC
0 3 5 10
Wavelength (λ in μm)

Figure 6. Emissivity measurements of the samples with an uncertainty less than 3%.

As mentioned before, the calibrations were done with oxidized samples. During machining there is no
oxidation. Then, the temperatures that are obtained when applying the calibration curve to the counts
obtained in the experiments are lower than in the reality. Therefore, in order to know the real temperatures,
we applied Planck’s law that enables to infer temperatures from thermal radiation.

Planck’s equation for a:


Surface with oxidation Surface with no oxidation
λ2 λ2
C1L C1L
Lλ ,oxi (λ , Toxi ) ≡ ∫ ε oxi (λ , Toxi ). Lnoxi (λ , Tnoxi ) ≡ ∫ ε noxi (λ , Tnoxi ).
⎛ C2
⎞ ⎛ λ .Tnoxi
C2

λ1
λ .⎜⎜ e
5 λ .Toxi
− 1⎟⎟ λ1

λ .⎜ e
5
− 1⎟⎟
⎝ ⎠ ⎝ ⎠

Lλ ,noxi (λ , Tnoxi ) ≡ Lλ ,oxi (λ , Toxi )

We used an iterative methodology to equal the radiance of the surface with oxidation to the surface with no
oxidation by changing the value of Toxi and giving the correspondent emissivity values to the chip and tool.

At each set of cutting conditions of 1 second of machining time, a thermal “movie” of approximately 300 to
400 frames of data was obtained. These data were converted to temperature using the appropriate
calibration curve and the iterative procedure (mentioned above) in order to observe if the steady state was
reached in each cutting condition. Once the steady state was found in all the cutting conditions, the same 50
frames were chosen for all the cutting conditions and compare between them.

These 50 frames (steady state situation) were used to produce a mean and standard deviation thermal maps
for each cutting condition. By dividing the standard deviation at each pixel by the square root of the number
of samples, uncertainty in the mean temperature at each pixel was generated. From these two thermal maps,
three different thermal profiles were extracted; I) along the rake face in the chip side (15 microns above the
interface), II) along the rake face in the tool side (15 microns below the interface of the interface) and III) flank
face to observe the maximum temperature.

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


ºC
600
CHIP
400

200
TOOL

WORKPIECE

15µm above tool-chip interface (I)


15µm below tool-chip interface (II)
Flank face (III) 100µm

Figure 6. Profiles extracted from the thermal maps.

On balance, in these particular measurements there were three primary sources of error:
I) The first type of error was due to experimental fluctuation. This could be because focus variations,
rapid fluctuations in local surface roughness, stray light from the chip and/or environment, etc. The
uncertainty in measurement of the mean temperature at each pixel was calculated from the 50
images by calculating the standard deviation at each pixel and dividing by the square root of the
number of samples.
II) The second type of error was due to calibration measurements. Numerical errors in the iterative
calculation of temperature, focal plane array non uniformity, different sensitivity, etc. The NUC and
sensitivity effect of the plane array was diminished by calibrating each pixel one by one. Thereby,
the error source was left to the mathematical approximations mainly.
III) The third type of error was due to shift in the mean surface emissivity. The work piece is exerted to
severe plastic deformation and this lead to change the angle of observation, roughness and maybe
the oxidation status. First, it is assumed that there is no time for oxidation because the chip is
approximately 1 millisecond in the field of view whereas it has been cited that it is required several
seconds to form a compact oxide layer ranging from 3 to 5 microns. This is the minimum thickness
in order to change significantly the emissivity. Secondly, it is obvious that since there is a plastic
deformation the surface roughness changes as it has been seen in some Hopkinson tests. The
uncertainty in emissivity has been determined as 10 percent for this case. On the other hand, the
tool is not subjected to any change in oxidation or observation angle. Moreover, in order to reduce
as much as possible the uncertainty of the surface roughness, it was polished with a silicon carbide
P180 grade to keep the roughness constant. Thus, the error that has been considered is 3 percent
that is the uncertainty of the emissivity measurements.

Uncertainties Workpiece Tool


Calibration Measurement System ±10 C ±10 C
Shift in Mean Surface emissivity ± 25 C ± 25 C
Experimental fluctuation ± 35 C ±10 C
Combined uncertainty ± 44 C ±28 C
Table 1. Temperature measurements uncertainty.

RESULTS
After running the experiments and applying the calibration curves in addition to Planck’s law, we obtained a
thermal map for each cutting condition and the corresponding uncertainty map. Representative thermal fields
and maps of the uncertainties in the mean temperature obtained from the experimental fluctuations are shown
in Figure 7. It is clear that the maximum temperatures are in the secondary zone and the tool is the hottest
element. Besides, the chip temperature gradually decreases starting at the rake face (tool face) until free
surface (top surface of the chip).

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


ºC 800 ºC
25 Chip
600 Chip 15
200 5 Tool
Tool
Workpiece Workpiece

100µm 100µm
a) b)

ºC 800 ºC
25 Chip
Chip
600 15
Tool

200 5 Tool
Workpiece
Workpiece

100µm 100µm

c) d)
Figure 7. Thermal maps of 4140 Standard temperature (a) and its uncertainty (b).and. 4140 MECAMAX® PLUS
-1
temperature (c) and its uncertainty (d) under the cutting speed of 200 m·min and uncut chip thickness of 0.2mm.

Note that the experimental fluctuations also appear to be highest at the tool-chip interface and that the
temperatures in cutting the standard grade are significantly higher over the whole field in the AISI 4140
Standard. Figure 8 makes this more quantitative showing maximum temperature in the chip and tool for
various parameters. From 100 m·min-1 to 300 m·min-1 the difference remains approximately in 40ºC for both
the chip and tool, but jumps to 95ºC (chip side) and 75ºC (tool side) at 400 m·min-1. Meanwhile, in the flank
face the differences range from 30 to 50 ºC. We have shown just one temperature for cutting speed of 500
m·min-1 to show the temperatures increase drastically, mainly, due to the tool wearing.

Machinability Effect
Machinability Effect
1100 1100

1000 1000
Temperature (ºC)
Temperature (ºC)

EstandarChip
900 900
EstandarTool EstandarFlank
PlusChip PlusFlank
800 PlusTool 800

700 700

600 600
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Cutting Speed (m.min-1) Cutting Speed (m.min-1)

Figure 8. Experimentally measured maximum tool temperature (solid line) and maximum chip temperature (dotted
line) versus cutting speed. The red is AISI 4140 Standard and the green is AISI 4140 MECAMAX® PLUS.

This differences in temperature are believed to be due to the different treatments or adding different elements
that AISI 4140 Plus has been subjected to enhance its machinability rate without changing the mechanical
properties or microstructure characteristic.

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


SUMMARY
On one hand, from a scientific point of view, it has been shown how IR cameras can provide accurate data
which help to understand more detail in the machining where a lot of complex thermo mechanical phenomena
are involved in a really short time and a small area. Furthermore, it can be said that in order to measure
temperatures properly with the infrared cameras, it is required an accurate samples’ surface characterization
and corresponding emissivity measurements.

On the other hand, from a manufacturing point of view, it has been observed that across a range of
parameters that enhanced machinability in AISI 4140 MECAMAX® PLUS, does lead to reduced temperatures
in the tool-chip region. Furthermore, the results indicate that in certain critical regions of the thermal field,
improved machinability correlates with reductions in temperature that exceed measurement uncertainties,
most notably in the tool where the measurement uncertainty is only ±28ºC. The greatest differences were
observed at the higher cutting speed of 400 m·min-1, which is consistent with the precipitation and consequent
deposition of the MnS layer on the rake face [6]. The method developed here may be a practical tool for the
scientific design of more machinable materials and reduce the need for tedious, time-consuming and
expensive than machinability tests (ISO 3685).

REFERENCES
[1] M. Eugene Merchant, 1998, “An interpretive look at 20th century research on modelling of machining”,
Proceeding of the CIRP international workshop on modelling of machining operations, Georgia Instit. of
technology. Atlanta.

[2] Gol’dshtein, Ya. E., Morozov, A.N., ‘Steels with selenium and calcium’, Metal Science and Heat Treatment
(English Translation of Metallovedenie i Termicheskaya Obrabotka Metallov), v 22, n 11-12, Nov-Dec, 1980, p
787-793.

[3] Grzesik, W. , ‘Experimental investigation of the cutting temperature when turning with coated indexable
insert’, International Journal of Machine Tools & Manufacture, v 39, n 3, Mar, 1999, p 355-369.

[4] Wanigarathne, P.C., Ee, K.C., Jawahir, I.S., ‘Near-dry machining for environmentally benign
manufacturing’, Design and Manufacture for Sustainable Development 2003, Design and Manufacture for
Sustainable Development 2003, 2003, p 39-48.

[5] Hamann, J.C., Grolleau,V., Le Maître, F. Machinability improvement of steels at high cutting speeds -
study of tool/work material interaction. Annals of the CIRP, Vol.45/1/1996, pp.87-92.

[6] Da Silva, M.B., Wallbank, J. (1999), "Cutting temperature: Prediction and measurement methods – a
review", Journal of Materials Processing Technology, Vol. 88 pp.195-202.

[7] Rumford, B. C. o., 1798, An Inquiry concerning the Source of the Heat which is excited by Friction,
presented at Philisophical Transactions of the Royal Society of London.

[8] Morgan, M.N. (2004), "Developments in temperature measurement and process monitoring systems for
grinding", paper presented at IGT@UWE Annual Seminar, Bristol, 18 May. Available also at
http://www.smartfibres.com/Smart_FBG_Sensors.htm (accessed 19 May 2004).

[9] Dewes (1999), "Temperature measurement when high speed machining hardened mould/die steel",
Journal of Materials Processing Technology, Vol. 92-93 pp.293-301.

[10] Yoshioka, H., Hashizume, H., Shinno, H., 2004, In-process microsensor for ultraprecision machining, IEE
Proceedings - Science, Measurement and Technology, 151/2:121-125.

[11] Kato, T., Fujii, H. (1996), "PVD film method for measuring the temperature distribution in cutting tools",
ASME Journal of Engineering for Industry, Vol. 118 pp.117-22.

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


[12] Reis, D. D., Abrão, 2005, The machining of aluminum alloy 6351, Journal of Engineering Manufacture,
219/27-33.

[13] Weinert, K., Tillmann, W., Hammer, N., Kempmann, C., Vogli, E., 2006, Tool coatings as thermocouple
for the monitoring of temperatures in turning processes, Advanced Engineering Materials, 8/10:1007-1010.

[14] Potdar, Y.K., Zehnder, A.T., 2004, ‘Temperature and Deformation Measurements in Trasient Metal
Cutting’, Experimental Mechanics, Vol. 44, No.1, Society for Experimental Mechanics

[15] Davies, M.A., Cooke, A.L., Larsen E.R., 2005, ‘High Bandwidth Thermal Microscopy of Machining
AISI1045 Steel’, 55th 2005 CIRP General Assembly, Turkey

[16] Pujana, J., Del Campo, L., Perez-Saez, R.B., Tello, M.J., Gallego, I., Arrazola, P.J., ‘Radiation
thermometry applied to temperature measurement in the cutting process’, Measurement Science and
Technology, v 18, n 11, Nov 1, 2007, p 3409-3416.

ACKNOWLEDGEMENTS
To the Basque and Spanish Governments for the financial support given to the projects: INCLUMEC (UE200703) and
MAQUIMODEL (DPI2006-15502-C02-01). Authors also thank the company Sidenor (Spain) for the technical and material
support given to this work. Thanks to the University of the Basque Country and University of North Carolina at Charlotte
fore emissivity measurements and infrared camera usage respectively.

ABOUT THE AUTHOR


Ivan Arriola Aldamiz is a PhD student from Mondragón University, Spain. He has been involved in high
performance machining for the last 4 years during his undergraduate project in the University of Kentucky
from 2003 to 2004, Masters in the University of Mondragon from 2004 to 2007 and in the University of North
of Carolina at Charlotte from 2007 to 2008. Currently, he is studying the effect of the machinability of steels on
machining temperatures at NIST (Gaithersburg).

InfraMation 2008 Proceedings ITC 126 A 2008-05-14


InfraMation 2008 Proceedings ITC 126 A 2008-05-14

You might also like