You are on page 1of 8

Journal of Materials Processing Technology 214 (2014) 12931300

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


journal homepage: www.elsevier.com/locate/jmatprotec

Cutting temperatures during hard turningMeasurements and effects


on white layer formation in AISI 52100
S.B. Hosseini a, , T. Beno b , U. Klement a , J. Kaminski a , K. Ryttberg c
a
Department of Materials and Manufacturing Technology, Chalmers University of Technology, SE-412 96 Gothenburg, Sweden
b
Department of Engineering Science, University West, SE-461 86 Trollhttan, Sweden
c
AB SKF, SE-415 50 Gothenburg, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: This paper concerns the temperature evolution during white layer formation induced by hard turning
Received 8 July 2013 of martensitic and bainitic hardened AISI 52100 steel, as well as the effects of cutting temperatures
Received in revised form 8 December 2013 and surface cooling rates on the microstructure and properties of the induced white layers. The cutting
Accepted 27 January 2014
temperatures were measured using a high speed two-colour pyrometer, equipped with an optical bre
Available online 4 February 2014
allowing for temperature measurements at the cutting edge. Depending on the machining conditions,
white layers were shown to have formed both above and well below the parent austenitic transformation
Keywords:
temperature, Ac1 , of about 750 C. Thus at least two different mechanisms, phase transformation above
White layer
Hard turning
the Ac1 (thermally) and severe plastic deformation below the Ac1 (mechanically), have been active during
Phase transformation white layer formation. In the case of the predominantly thermally induced white layers, the cutting
Intense plastic deformation temperatures were above 900 C, while for the predominantly mechanically induced white layers the
Cutting temperatures cutting temperatures were approximately 550 C. The surface cooling rates during hard turning were
Surface integrity shown to be as high as 104 105 C/s for cutting speeds between 30 and 260 m/min independent of whether
the studied microstructure was martensitic or bainitic. Adding the results from the cutting temperature
measurements to previous results on the retained austenite contents and residual stresses of the white
layers, it can be summarised that thermally induced white layers contain signicantly higher amounts
of retained austenite compared to the unaffected material and display high tensile residual stresses. On
the contrary, in the case of white layers formed mainly due to severe plastic deformation, no retained
austenite could be measured and the surface and subsurface residual stresses were compressive.
2014 Elsevier B.V. All rights reserved.

1. Introduction Choudhury (2012), the major reasons are related to the tool wear,
which will result in larger dimensional variations (geometric form
Historically, grinding has been used for machining of hard- and surface roughness), alterations of the surface and subsurface
ened steel parts. However, in the past few decades hard turning microstructures and generation of high tensile residual stresses.
has gained increased interest from both the scientic community All these properties have a negative inuence on the fatigue perfor-
as well as the industry. Advantages of hard turning, if compared mance. For example, when studying the relationship between the
to grinding, are e.g. dry cutting possibilities, turning of complex residual stresses, white layers and the fatigue life of hard turned
geometries and the possibility of performing rough and nish tur- surfaces of AISI 52100 steel, Guo et al. (2010) concluded that white
ning in one clamping. One example is given by Tnshoff et al. (2000) layers could reduce the fatigue life by as much as 8 times com-
who compared grinding and hard turning of a hydraulic component pared to surfaces free from white layers. On the contrary, Smith
with respect to the surface quality, machining time and the total et al. (2007) concluded that there was not any conclusive evidence
cost. They concluded that a comparable surface quality could be suggesting that the hard turned induced white layers (2 m) had a
achieved by hard turning, at the same time as the total machin- negative inuence on the axial fatigue life of AISI 52100 steel.
ing time was reduced by 60%. However, even though there is a The developed microstructure is referred to consist of white and
strong interest in adapting hard turning, the process is still not as dark layers, and displays properties which are signicantly different
widely used in industry as expected. As discussed by Bartarya and from the bulk material, such as differing hardness and rened grain
structure. The nomenclature has developed as this microstructure,
when polished and etched, appears whiter (white) and/or darker
Corresponding author. Tel.: +46 31 337 2854. (dark), compared to the unaffected material when viewed in a light
E-mail addresses: seyed.hosseini@chalmers.se, seyed.b.hosseini@gmail.com optical microscope, see Fig. 1. The thicknesses of the respective lay-
(S.B. Hosseini). ers as generated in hard turning are typically up to 5 m for white

0924-0136/$ see front matter 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jmatprotec.2014.01.016
1294 S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300

by the authors. Mller et al. (2004) used a two-colour pyrometer to


determine both chip and work piece surface temperatures for AISI
1045 steel. The work piece surface temperatures were reported to
increase with the cutting speed and reached a maximum tempera-
ture of 700 C. More recently, Han et al. (2008) measured the work
piece surface temperatures of AISI 1045 steel by using thermocou-
ples located at the tool/work piece interface in order to validate
modelled temperatures. Despite deviations of up to 32% between
measured and calculated temperatures, the model was considered
to be accurate enough for predicting the surface temperatures.
Formation of white and dark layers during hard turning of
AISI 52100 steel with martensitic and bainitic microstructures has
previously been studied by Hosseini et al. (2012). The inuences
of cutting speed and tool wear on the formation of white and
dark layers as well as the characteristics of the resulting micro-
structures were then thoroughly studied. The authors concluded
Fig. 1. White and dark layers induced by hard turning. (vc = 110 m/min and that depending on the cutting conditions, the white layers had been
Vb = 0.3 mm). predominantly thermally or mechanically induced. For the pre-
dominantly thermally induced white layers, the authors estimated
the cutting temperatures by analysing the microstructural con-
layers and up to 15 m for dark layers, even though white layers as stituents in relation to a timetemperature austenitisation diagram
thick as 12 m have been reported in the literature. (TTA). However, no experimental evidence for the temperatures
As discussed by Grifths (1987) there are basically three main was provided. The aim of the present study was to measure the
mechanisms suggested for the formation of white layers in vari- cutting temperatures during hard turning of AISI 52100 steel under
ous manufacturing processes: (i) phase transformation due to rapid comparable cutting conditions using a two-colour pyrometer, and
heating followed by a subsequent rapid cooling, (ii) severe plastic to correlate the measured temperatures with the previous results
deformation resulting in a ne grain structure, and (iii) chemical on the white layer microstructure constituents and the resid-
surface reaction with the environment. Even though few attempts ual stresses. Before the hard turning experiments, the accuracy
to measure the temperatures during hard turning have been pre- of the two-colour pyrometer was investigated by Hosseini et al.
sented, it is generally believed that white layer formation is mainly (2013), and due to the large temperature deviations between the
thermally driven, in other words the machined surface is rapidly two-colour pyrometer and the thermocouple readings, an approx-
heated above the austenitic phase transformation temperature imation function was developed for the studied materials. The
(Ac1 ) followed by a rapid quenching. To get an indication of whether developed approximation function has been applied in the current
the machined surface has been heated above Ac1 , the changes in the study to correct the measured temperatures.
retained austenite content are often analysed by means of X-ray
diffraction. For example, Chou and Evans (1999) measured a three-
2. Experimental procedure
fold increase of the retained austenite content in the white layers
compared to the unaffected material when studying hard turning of
2.1. Work piece material
SAE 52100 steel, which made the authors conclude that the forma-
tion had been mainly thermally activated (rapid heating and rapid
The material chosen for this study was AISI 52100 steel with
cooling). There are also some works highlighting the importance
the nominal chemical composition as shown in Table 1. Since
of severe plastic deformation during white layer formation. After
the studied material is often heat treated to have either bainitic
observing a fourfold reduction in the retained austenite content in
or martensitic structure, both structures were investigated. Prior
the surface area after hard turning of AISI 52100 steel at low cut-
to the nal hardening treatment, the microstructure of the steel
ting speed, Ramesh et al. (2005) concluded that the subsequently
consisted of evenly distributed spheroidised (Fe,Cr)3 C carbides in
induced white layers were formed due to severe plastic deforma-
a ferritic matrix with a hardness of 200 HV30. After the nal heat
tion. Furthermore, based on the similarities of the microstructure
treatment, the hardnesses of the studied martensitic and bainitic
and crystallography of the white layers induced by hard turning
microstructures were 61 and 60 HRC, respectively.
and adiabatic shear bands in general, Barry and Byrne (2002) pro-
posed that the main mechanism causing white layer formation
was dynamic recovery. However, the authors also emphasised the 2.2. Hard turning
importance of dynamic recrystallisation for white layers generated
during hard turning with worn cutting tools. To ensure the formation of white layers during hard turning,
To clarify, if hard turned induced white layers are predominantly the cutting speed and the tool ank wear were kept at the same
thermally (temperature > Ac1 ) or mechanically (temperature < Ac1 ) levels as used in the previous investigation made by the authors
activated, temperature measurements are necessary. However, (Hosseini et al., 2012), where three different cutting speeds (30, 110
there are only a few attempts presented in the literature where and 260 m/min) and two levels of tool ank wear (0 and 0.2 mm)
both the work piece and the cutting tool temperatures have been were studied using a CBN insert of grade BNX10. However, in the
measured during hard turning. For example, Ueda et al. (1999) used present investigation, a CBN insert of grade BNC200 was chosen
a two-colour pyrometer for measuring the cutting edge tempera- due to its benecial interrupted cutting properties. The insert had
tures during internal turning of AISI 52100 steel. They concluded a nose radius of 0.8 mm, a chamfer land and angle of 0.10 mm and
that by increasing the cutting speed from 100 m/min to 300 m/min, 15 , respectively and a clearance angle of 7 . The feed rate and depth
the temperatures increased from 800 C to 950 C. However, it of cut were also chosen according to the above-mentioned studies;
was not clear whether the stated temperatures were averaged or 0.08 mm/rev and 0.08 mm, respectively, and the cutting was per-
maximum temperatures. Also, no clear relation between the tool formed in dry conditions. To prevent premature tool failure during
conditions and the measured cutting temperatures was provided the temperature measurements when utilising worn tools, the tools
S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300 1295

Table 1
Nominal chemical composition (wt.%) of the work material, Fe-bal.

Material C Mn Cr Si S P Ni

AISI 52100 0.951.05 0.200.45 1.301.65 0.200.35 0.015 max 0.027 max 0.30 max

were pre-worn on specimens with comparable geometry, micro- beam was sent through the bre. With the help of the micrometre
structure and hardness. Once a tool ank wear of 0.18 mm was screws the spot was adjusted to view the tip of the cutting tool. The
measured, the cutting edge or surface temperatures were recorded functionality and the accuracy of the system were tested by several
using either test samples with pre-machined slot or tools with the pre-runs.
through-hole; see below.

2.3.2. Cooling rate calculations


2.3. Temperature measurements
To determine the surface cooling rates during hard turning, tem-
peratures from at least two well dened positions are required. In
Temperatures were measured by means of a two-colour pyrom-
the current study, besides the cutting tool temperatures, the sur-
eter (described in Section 2.4) both at the edge of the cutting tool
face temperatures were also measured at a pre-dened distance
and at the work piece surface at a short distance beneath the cut-
beneath the cutting edge. To be able to measure the surface tem-
ting edge. The rst measurement corresponded to the actual cutting
peratures close enough to the cutting edge, a through-hole with
temperature while the second type of measurement allowed for
a diameter of about 0.7 mm was made in the cutting tool 1.5 mm
calculations of the cooling rate. The experimental setups for the
beneath the cutting edge using electric discharge machining; see
respective types are presented in Fig. 2 and described in more detail
Fig. 2(c). To verify that the through-hole did not cause early or
below. For both types: cutting tool temperatures and cooling rate
unexpected tool failure, several pre-tests were conducted before
calculations, one temperature measurement per cutting condition
the actual experiments.
and microstructure was performed, resulting in a total of 27 tests.
Depending on the cutting speed, each experimental run was carried
out for 1050 s during which temperatures were recorded. In the 2.4. Two-colour pyrometer
case of the cutting edge temperature measurements this resulted
in about 30300 temperature proles per experimental condition The contactless two-colour pyrometer used in the current inves-
since one prole was recorded per revolution (passing of the slot). tigation was developed by Mller and Renz (2001). The basic
For the work piece surface temperatures used for the cooling rate concept of a pyrometer is to detect the infrared radiation emitted
calculations, a continuous signal was recorded. This was due to the from a target and the temperature can be derived by knowing the
fact that the bre was pointing towards the newly machined sur- emittance value,  , of the target surface at a specic wavelength.
face through the cutting tool, which means that the entire cutting However, when using a two-colour pyrometer, the temperatures
time could be used for analysis. The temperatures provided in Sec- can be determined without knowledge of the emittance values by
tions 3 and 4 are the averaged temperature values and the error just assuming grey-body condition, i.e. assuming the emittance val-
bars show the measured maximum and minimum temperatures. ues to be identical for the two wavelengths used (1 = 2 at 1 and
2 ). But even if the wavelengths (1 and 2 ) are chosen close to
2.3.1. Cutting temperature measurements each other the spectral emittance values for most metallic sur-
To carry out the temperature measurements at the edge of the faces differ substantially. This was for example demonstrated by
cutting tool, which corresponds to the actual cutting temperature, Kobayashi et al. (1999) in their study of the changes in the nor-
a slot with a length of 25 mm and a width of 0.4 mm (see Fig. 2(a)) mal spectral emittance values at high temperatures for numerous
was machined in each test ring by the use of wire-electric discharge pure metals and alloys. Recently, Hosseini et al. (2013) showed that
machining. During hard turning, the temperatures at the edge were for AISI 52100 steel hard turned with either fresh or worn cutting
measured each time the slot passed the cutting tool and the optical tools the applied two-colour pyrometer measured temperatures
bre. The bre was positioned inside the rings via a specially man- that were 1220% lower than the ones measured by thermocou-
ufactured tool holder as shown in Fig. 2(b). For each revolution, the ples that were considered to be correct target temperatures.
emitted radiation from the cutting tool was recorded and converted When measuring on the cutting tools, the two-colour pyrometer
into a unique temperature prole by the pyrometer software. To recorded 13% lower temperatures compared to the true target
ensure that correct temperatures were measured for each revolu- temperatures. Based on the ndings, three unique equations were
tion, precise positioning of the bre viewing through the slot was determined; one for each of the machined surfaces hard turned
important. To mount the bre in the right position, a red diode light with either fresh or worn cutting tools (Eqs. (1) and (2)) and one

Fig. 2. Illustration of (a) test sample with a slot; (b) tool holder used for cutting tool temperature measurements: (1) guide screw for the bre and (2) m-precision screws
for alignment of the bre; and (c) through-hole in the cutting tools used for work piece surface temperature measurements.
1296 S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300

for the cutting tool material (Eq. (3)), with which the temperatures
as obtained from the two-colour pyrometer were corrected:

TCorr = 1.27 TMeas 100 (1)

TCorr = 1.37 TMeas 68 (2)

TCorr = 1.17 TMeas 40 (3)

Here TCorr is the calculated target temperature in degrees Celsius


and TMeas is the temperature measured by the two-colour pyrom-
eter in degrees Celsius. These equations have been applied in the
current study to correct the temperatures as measured by the two-
colour pyrometer.
Fig. 3. Cutting edge temperatures versus cutting speed for the studied micro-
structures and tool conditions (M: martensitic and B: bainitic microstructures). The
3. Results
Ac1 line represents the austenitisation temperature for the material.

3.1. Cutting temperatures


At the highest cutting speed, the worn cutting tools resulted in a
The averaged cutting temperatures for both fresh and worn temperature increase of 50 ; from 850 C to 900 C.
cutting tools are provided in Fig. 3. In general, increased cutting
speeds and/or hard turning with worn instead of fresh cutting 3.2. Cooling rate calculations
tools resulted in increased cutting temperatures, while the micro-
structure of the test samples (bainite or martensite) did not have The cooling rates were determined by considering the differ-
a signicant effect on the temperature development. For exam- ence in temperatures measured at the cutting edge and at the work
ple, when hard turning samples with a martensitic structure using piece surface at a pre-dened distance beneath the cutting edge.
fresh cutting tools, increasing the cutting speed from 30 m/min to The surface cooling rates for each cutting condition and material
110 m/min resulted in temperatures changing from about 510 C are shown in Fig. 4(a)(c). Position A in each graph represents the
to about 810 C, while at the highest cutting speed (260 m/min) the averaged cutting temperatures measured at the edge of the cutting
measured temperatures were about 850 C. Changing from fresh tool, while position B corresponds to the averaged temperatures
to worn cutting tools always resulted in increased cutting temper- measured 1.15 mm beneath the cutting edge. As can be seen in
atures where the largest temperature increase was seen for the the graphs in Fig. 4, the lowest cooling rate was seen for surfaces
surfaces hard turned at 110 m/min, where the worn cutting tool turned at 30 m/min using worn tools and the highest for surfaces
made the cutting temperatures for martensitic samples rise from hard turned at 260 m/min using fresh tools.
810 C to 880 C. At the lowest cutting speed, 30 m/min, the tem- No signicant differences with respect to the cooling rates were
perature increase was not as large; 30 for both microstructures. seen between bainite and martensite. To verify the calculated

Fig. 4. Calculated cooling rates during hard turning with fresh and worn cutting tools: (a) 30 m/min, (b) 110 m/min and (c) 260 m/min. Symbols in the graphs represent
() worn and () fresh tools for martensitic structures and () worn and () fresh tools for bainitic microstructures, respectively. The () shows the work piece surface
temperature at 0.7 mm during hard turning with fresh cutting tools for samples with martensitic microstructures.
S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300 1297

cooling rates, additional work piece surface temperature measure- stress 1.3 GPa) by using ClausiusClayperon equation when tak-
ments at a shorter distance, 0.7 mm, beneath the cutting edge were ing into account the pressure effect for SAE 52100 steel. By adjusting
carried out. As shown by position C in Fig. 4(a)(c) the calculated the timetemperature-austenitisation diagram with the above
cooling rates at 0.7 mm conrmed the high cooling rates in hard mentioned parameters, the temperature differences between esti-
turning. mations and measurements would be signicantly lowered.

4.2. Cooling rate calculations


4. Discussion
In the current investigation, it was experimentally shown that
For each cutting condition the averaged temperature was pre-
even though cutting speeds as low as 30 m/min were used dur-
sented together with a temperature range representing the lowest
ing hard turning, the cooling rates were in the order of 104 C/s.
and highest measured temperatures for the specic condition,
By increasing the cutting speed to 260 m/min, the cooling rate was
which reects the accuracy of the two-colour pyrometer as well
increased by one order of magnitude to 105 C/s, which means
as the experimental setup. As shown in Figs. 3 and 4, the tem-
that the surface cooling rates during hard turning at the stud-
perature range was between 70 C and 170 C. Part of this range
ied cutting conditions were between 104 C/s and 105 C/s. This is
was considered to be from the two-colour pyrometer (accuracy
consistent with the results presented by Chou and Evans (1999),
and response time) and some part from the experimental setup.
who used Jeagers moving heat source theory to simulate the tem-
When studying the applicability and accuracy of the two-colour
perature elds generated in hard turned surfaces, from which the
pyrometer, Hosseini et al. (2013) observed that at an object tem-
cooling rate was calculated to be 4 104 C/s. Also comparable
perature of 570 C, the temperature range was not larger than 10 C
results were obtained by Uhlmann et al. (2013), who used a ther-
after three consecutive measurements. Also regarding the response
mography camera Jade Mwir II to investigate the cooling rates
time, it can be concluded from the work undertaken by Davies
during hard turning of AISI 52100 steel. At cutting speeds between
et al. (2007) after reviewing several techniques for measuring the
100 m/min and 200 m/min, the authors calculated the cooling rate
cutting temperatures that, in comparison to other techniques, two-
to be 8 104 1.5 105 C/s. In the current study, the tool wear was
colour pyrometers are superior and should be able to monitor rapid
shown to inuence the cooling rate; lower cooling rates were mea-
thermal changes. Hence, the measured range in the current investi-
sured for surfaces hard turned with worn cutting tools compared to
gation was not considered to be caused primarily by the two-colour
surfaces turned with fresh tools. This could probably be explained
pyrometer. On the other hand, from an experimental point of view,
by the differences in the contact times between the cutting tool and
it is possible that the measured temperatures are inuenced by e.g.
the work piece. For example, when the tool ank wear increased
particles or fragments of the produced chips, which could have a
from 0.04 mm to 0.2 mm during hard turning at 110 m/min, the
higher or a lower temperature than the area of interest.
contact time increased from 21 s to 110 s, which will result in a
larger heated volume and thereby lower cooling rates.
4.1. Cutting temperatures To obtain a better understanding of the formation mechanisms
of the white layer and the corresponding mechanical properties
In this investigation, the cutting tool temperatures were mea- as induced by hard turning, the temperatures as measured here
sured when the tool entered the slot during machining. By are discussed in combination with other surface responses in the
assuming that the temperature at the edge of the cutting tool is following section.
equal to the temperature at the work piece surface when the cut-
ting tool enters the slot, the temperatures at the machined surfaces 4.3. Relation between temperature, residual stress, retained
could be obtained. A similar approach for measuring the cutting tool austenite content and white layer
edge temperatures as described in the present investigation was
previously employed by Ueda et al. (1999) and more recently by Detailed knowledge of the cutting temperatures as well as the
Tanaka et al. (2009) who obtained comparable cutting edge temper- cooling rates present during white layer formation is fundamental
atures during hard turning of SAE 52100 and SCM 415 steel. Tanaka to understanding the evolution of the microstructure. For example,
et al. (2009) reported a temperature increase of 65 when the knowledge of the cutting temperatures helps to understand the
tool wear was increased from 0.0 mm to 0.1 mm for cutting speeds changes found in the retained austenite content, residual stresses
between 100 m/min and 150 m/min. Ueda et al. (1999) reported and the white and dark layer thicknesses, which are the main
that the cutting temperatures increased by some 120 when the parameters of importance for the life of a component in service.
cutting speed was increased from 100 m/min to 250 m/min during Moreover, since residual stresses are a surface response that can be
hard turning of SAE 52100 steel. By using a timetemperature- easily measured, they are often addressed to conclude whether the
austenitisation diagram, assuming a heating rate of 105 C/s and machined surfaces have been predominantly thermally or mechan-
considering the microstructure constituents, Hosseini et al. (2012) ically affected. For example, Jacobson et al. (2002) concluded that
estimated the temperature for surfaces machined at 110 m/min higher cutting speeds led to higher surface tensile residual stresses
and 260 m/min with worn cutting tools would reach 1200 C. In due to the increased cutting temperatures, while surface compres-
the current investigation, the temperature corresponding to the sive residual stresses were dominating at lower cutting speeds.
same cutting conditions was measured at 900 C, i.e. 300 below The relationship between white and dark layer thicknesses,
the estimated temperature. The main reason for the discrepancy retained austenite contents, and residual stresses for surfaces
between calculated and measured temperatures is considered to machined with comparable cutting conditions (cutting speed, tool
be due to the simplications made for the temperature estima- ank wear, feed rate, depth of cut, nose radius, hard turning
tions. For example, the effects of parameters like contact pressure machine, material batch) as used here were previously studied
(stress), rapid heating, plastic strain, latent heat during phase by the authors. Based on the changes in the surface responses for
transformation and changes in chemical composition caused by the different cutting conditions resulting in white layers, Hosseini
diffusion were not considered. In order to make the estimations et al. (2012) concluded whether the white layer formation had been
more accurate, those parameters should be taken into account. predominantly thermally or mechanically initiated. Consequently,
For example, Ramesh and Melkote (2008) estimated the phase the temperatures as measured in the present study provide valu-
transformation temperature would drop by at least 100 C (shear able information as to the conclusion of whether white layers are
1298 S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300

Fig. 5. Correlation between (a) cutting temperatures and residual stresses, (b) retained austenite content and residual stresses, and (c) retained austenite content and cutting
temperatures for the studied cutting conditions. The dots in each graph correspond to the investigated cutting conditions. The dark grey-coloured areas in the graphs represent
the 95% condence interval for the regression model and the bright grey-coloured areas are the 95% prediction intervals for a new single observation.
The retained austenite contents and residual stresses taken from Hosseini et al. (2012).

formed due to rapid heating above the austenitisation temperature the work material and the cutting tool, since a larger tool ank wear
(thermally) or due to severe plastic deformation (mechanically) allows the material to be austenitised for a longer time (between 4
at temperatures well below the Ac1 temperature. For example, and 5 times), resulting in both higher retained austenite contents
when studying the correlation between different surface responses and thicker white layers if compared to when using fresh cutting
(Fig. 5(a)), it can be seen that e.g. when the cutting temperatures tools. This is conrmed in Fig. 6(b), where it is seen that for sur-
increase the surface residual stresses also tend to increase and faces machined with fresh tools only discontinuous white layers
eventually shift from compressive to tensile (R2 = 0.83). Further- are found, and only for the highest cutting speed, while for the sur-
more, as shown in Fig. 5(b) a good correlation (R2 = 0.89) between faces machined with worn tools at the two highest cutting speeds
the surface retained austenite content and the surface residual the white layers are continuous and thicker. However, even though
stresses can also be seen, i.e. if the retained austenite content comparable temperatures were recorded at 110 and 260 m/min
increases, so do the surface residual stresses (tensile). Finally, as when using worn tools, the white layer thicknesses were not equal;
shown in Fig. 5(c), good correlation between the changes in the 1.5 m and 3 m, respectively. This could once again be explained
retained austenite content and the temperatures was also observed, by the difference in contact time between tool and work piece
R2 = 0.78. Thus, from this regression analyses it can be concluded surface. When machining at 110 m/min instead of 260 m/min, the
that even though the residual stresses cannot solely provide any increased contact duration will allow for a larger material volume
detailed information on the cutting temperatures, if combined with to be heated above Ac1 , which should result in a thicker white layer.
the retained austenite content they can be used with condence for However, even though a thicker white layer is initially created, part
determining whether the detected white layers have been ther- of this layer might be re-tempered resulting in the nal white layer
mally or mechanically induced. thickness being less than for the surfaces hard turned at higher cut-
Further comparisons of the retained austenite contents, the ting speeds. This should then be reected in signicantly larger dark
white and dark layer thicknesses and the cutting temperatures layers for surfaces machined at 110 m/min compared to 260 m/min.
are provided in Fig. 6. As shown in Fig. 6(a), cutting temperatures As seen in Fig. 6(b), this was also the case since the dark layer thick-
well above the Ac1 , generally resulted in increased retained austen- nesses for surfaces hard turned at 110 and 260 m/min were 10 m
ite contents. For surfaces machined using worn cutting tools the and 5 m, respectively.
increase was large, from below 1% to about 8%, while for surfaces For cutting temperatures 200 C below Ac1 , see Fig. 6(a), the
machined with fresh cutting tools it was much less. This was con- retained austenite contents had not increased but on the con-
sidered to be mainly due to the difference in contact times between trary decreased to undetectable levels. A reduction in the retained

Fig. 6. (a) The retained austenite contents in the white layers formed either on originally a martensitic or a bainitic microstructure versus the cutting temperatures (lines
A and B is the retained austenite content in the as received microstructures, bainitic and martensitic, respectively). The measurement accuracy for the retained austenite
contents was estimated to be 2 vol.% and the temperature bars in (a) and (b) are the measured temperature range. (b) White and dark layer thicknesses versus the cutting
temperatures. The white layer thickness variation for vc = 30 and 110 was 0.5 m and for vc = 260 it was estimated to be 1.0 m. The retained austenite contents and
white/dark layer thicknesses are taken from Hosseini et al. (2012). (vc in m/min).
S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300 1299

Fig. 7. Schematic illustration of the interaction between the cutting tool and work piece material for (a) new and (b) worn cutting tools. Compare the difference in the tertiary
shear zone (A A*) in (a) and (b). The tool crater wear is not included in the gures.

austenite content can be explained by the two following mecha- microstructure being either bainite or martensite did not result in
nisms: (i) decomposition of the austenite into ferrite and cementite signicantly differing cutting temperatures. White layers contain-
due to the generated heat, which can be compared to tempering or ing up to 10 (martensite) and 15 (bainite) times higher retained
(ii) strain-induced martensitic transformation causing the austen- austenite contents compared to the unaffected microstructure
ite to transform into martensite during plastic deformation. Since and displaying high surface residual stresses had been formed at
both of these mechanisms occur simultaneously during hard tur- temperatures between 840 C and 920 C. On the contrary, when
ning, it is most likely that a reduction in the retained austenite no retained austenite content could be detected and the surface
content is due to a combination of the two mechanisms. Com- as well as the subsurface residual stresses were compressive, the
parable results have been reported by Ramesh et al. (2005) in white layers had been formed well below the austenitic phase
their study of hard turning of SAE 52100 steel. When machining transformation temperature, at a temperature of 550 C.
at 91 m/min with fresh cutting tools, the retained austenite con- With respect to the cooling rates in hard turning, it was found
tent was reduced from 20 vol.% to 5 vol.%, which led the authors that independent of the starting microstructure and cutting con-
to conclude that the white layer formation was mainly caused ditions, the cooling rates were between 104 C/s and 105 C/s.
by severe plastic deformation. As seen in Fig. 6(b), white layers Increased cutting speeds resulted in higher cooling rates while
can also be formed at temperatures well below Ac1 which is then increased tool ank wear resulted in lower cooling rates, which
mainly caused by severe plastic deformation (mechanically). To was governed by the increased contact length between the cutting
be able to create predominantly mechanically induced white lay- insert and the work material.
ers, the tool wear land (A A* in Fig. 7) plays an important role as
the temperatures are not high enough to cause re-austenitisation.
However, although the cutting temperatures are below Ac1 , the Acknowledgements
thermal effects will cause a softening of the work piece material
and as the tool ank wear land (distance A A* in Fig. 7) slides The Area of Advanced Production at Chalmers University of Tech-
against the work piece surface the severe plastic deformation that nology is acknowledged for nancial support. The authors also
takes place at the surface will result in the formation of white lay- acknowledge The Production Technology Centre at University West
ers. In the case of worn cutting tools, this distance is signicantly for their contribution and support during the usage of the two-
increased, e.g. by a factor of 4 when the ank wear increases from colour pyrometer. Mr. S. Johansson, Dr. M. Escursell and Mr. U.
0.05 to 0.2 mm. As a consequence of the increased contact length Hulling are all greatly acknowledged for their support throughout
(A A* in Fig. 7) the surface will experience a higher degree of the entire work as regards for the temperature measurements and
plastic deformation resulting in the formation of white layers due design of the tool-xture.
to severe plastic deformation. This effect, of severe plastic defor-
mation can also be observed when studying the residual stress References
proles. For example, when Thiele and Melkote (2000) studied
the effect of tool edge geometry on the work piece subsurface Barry, J., Byrne, G., 2002. TEM study on the surface white layer in two turned hard-
deformation on AISI 52100 steel, the authors concluded that large ened steels. Mater. Sci. Eng. A 325, 356364.
Bartarya, G., Choudhury, S.K., 2012. State of the art in hard turning. Int. J. Mach. Tools
edge honed tools (comparable to worn tools) produced measurable Manuf. 53, 114.
subsurface plastic ow. On the contrary, when using a sharp cut- Chou, Y.K., Evans, C.J., 1999. White layers and thermal modeling of hard turned
ting tool (comparable to fresh tools), no subsurface ow could be surfaces. Int. J. Mach. Tools Manuf. 39, 18631881.
Davies, M.A., Ueda, T., MSaoubi, R., Mullany, B., Cooke, A.L., 2007. On the measure-
observed. ment of temperature in material removal processes. Ann. CIRP 56 (2), 581
604.
Grifths, B.J., 1987. Mechanisms of white layer generation with reference to machin-
5. Conclusions ing and deformation processes. J. Tribol. 109, 525530.
Guo, Y.B., Warren, A.W., Hashimoto, F., 2010. The basic relationships between resid-
In this investigation the cutting tool temperatures and work ual stress, white layer, and fatigue life of hard turned and ground surfaces in
rolling contact. CIRP J. Manuf. Sci. Technol. 2, 129134.
piece surface cooling rates were determined during white layer Han, S., Melkote, S.N., Haluska, M.S., Watkins, T.R., 2008. White layer formation due
formation induced by hard turning of AISI 52100 steel. By use of to phase transformation in orthogonal machining of AISI 1045 annealed steel.
a two-colour pyrometer, the cutting temperatures were measured Mater. Sci. Eng. A 488, 195204.
Hosseini, S.B., Beno, T., Johansson, S., Klement, U., Kaminski, J., Ryttberg, K., 2013.
and found to increase from 510 C to 850 C and from 540 C to A methodology for temperature correction when using two-color pyrom-
920 C, respectively, when the cutting speed was increased from 30 eters compensation for surface topography and material. Exp. Mechan.,
to 260 m/min for fresh and worn cutting tools. http://dx.doi.org/10.1007/S11340-013-9805-7 (in press).
Hosseini, S.B., Ryttberg, K., Kaminski, J., Klement, U., 2012. Characterization of the
It was shown that white layers induced by hard turning Surface Integrity induced by Hard Turning of Bainitic and Martensitic AISI 52100
can occur both below and above the austenitic phase transfor- Steel. In: Procedia CIRP, vol. 1, 5th CIRP Conference on High Performance Cutting
mation temperature, Ac1 . For comparable hardness values, the Suppl, pp. 494499.
1300 S.B. Hosseini et al. / Journal of Materials Processing Technology 214 (2014) 12931300

Jacobson, M., Dahlman, P., Gunnberg, F., 2002. Cutting speed inuence on surface Smith, S., Melkote, S.N., Lara-Curzio, E., Watkins, T.R., Allard, L., Riester, L., 2007.
integrity of hard turned bainitic steel. J. Mater. Process. Technol. 128, 318323. Effect of surface integrity of hard turned AISI 52100 steel on fatigue performance.
Kobayashi, M., Ono, A., Otsuki, M., Sakate, H., Sakuma, F., 1999. A database of normal Mater. Sci. Eng. A 459, 337346.
spectral emissivities of metals at high temperatures. Int. J. Themophys. 20 (1), Tanaka, R., Motishita, H., Lin, Y., Hosokawa, A., Ueda, T., Furumoto, T., 2009. Cutting
299308. tool edge temperature in nish hard turning of case hardened steel. Key Eng.
Mller, B., Renz, U., 2001. Development of a fast ber-optic two-color pyrometer for Mater. 407 (408), 268272.
the temperature measurement of surfaces with varying emissivities. Rev. Sci. Thiele, J.D., Melkote, S.N., 2000. Effect of tool edge geometry on workpiece subsurface
Instrum., AIP 72 (8), 33663374. deformation and through-thickness residual stresses for hard turning of AISI
Mller, B., Renz, U., Hoppe, S., Klocke, F., 2004. Radiation thermometry at a high- 52100 steel. J. Manuf. Processes Vol. 2 (4), 270276.
speed turning process. J. Manuf. Sci. Eng. 126, 488495. Tnshoff, H.K., Arendt, C., Ben Amor, R., 2000. Cutting of hardened steels. Ann. CIRP
Ramesh, A., Melkote, S.N., 2008. Modeling of white layer formation under thermally 49 (2), 547566.
dominant conditions in orthogonal machining of hardened AISI 52100 steel. Int. Ueda, T., Al Huda, M., Yamada, K., Nakayama, K., 1999. Temperature measurement
J. Mach. Tools Manuf. 48, 402414. of CBN tool in turning of high hardness steel. Ann. CIRP 48 (1), 6366.
Ramesh, A., Melkote, S.N., Allard, L.F., Riester, L., Watkins, T.R., 2005. Analysis of Uhlmann, E., Mahnken, R., Ivanov, I.M., Cheng, C., 2013. FEM modeling of hard turning
white layers formed in hard turning of AISI 52100 steel. Mater. Sci. Eng. A 390, with consideration of viscoplastic asymmetry and phase transformation. J. Mach.
8897. Eng. 13 (1), 8092.

You might also like