Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317

A computational approach to evaluate temperature and heat


partition in machining with multilayer coated tools
W. Grzesik ∗, P. Nieslony
Department of Manufacturing Engineering and Production Automation, Technical University of Opole, P.O. Box 321, Opole 45-271, Poland

Received 19 May 2003; accepted 10 June 2003

Abstract

In this paper, analytical models for estimating the interface temperature and heat partition to the chip in continuous dry machining
of steels with flat-faced tools treated with multilayer coatings are presented. The database for modeling includes changes in the
thermal properties of both workpiece and substrate/coating materials and the Peclet and Fourier numbers occurring at actual interface
temperatures. Process outputs involve the average tool–chip interface temperature, the tool–chip contact length, the friction energy
and the heat balance between the moving chip and stationary tool. It was found that the heat partition coefficient varies significantly
from 0.65 to 0.8 when using multilayer coated tools, and changes from 0.5 to 0.6 for uncoated carbide tools. This implies that the
use of multilayer coated tools causes about 30% more heat generated due to friction to be transferred into the moving chip. In
general, both power and linear models can be used to estimate the interface temperature.
 2003 Elsevier Ltd. All rights reserved.

Keywords: Machining; Steels; Multilayer coatings; Heat partition

1. Introduction interface are sensitive to changes in the thermophysical


properties of the workpiece and coating materials,
In metal cutting operations, temperature develops at including the thermal conductivity, thermal diffusivity
the tool–chip interface, and heat transferred from this and heat transfer coefficient [5].
zone is a crucial factor, determining such key process It is well known that temperature rise in metal cutting
issues as tool wear and its life, and surface integrity. is caused by two principle heat sources—the first
These influences are predominant in dry machining at resulting from plastic deformation developed at the pri-
high speed, which becomes an economical and ecologi- mary shear plane (the shear zone heat source) and the
cal imperative. While heat generation depends mainly on second due to friction at the tool–chip interface (the fric-
process parameters and the machinability rate of the tional heat source). There are many viable analytical and
work material, the thermophysical properties of the cut- computational/simulation methods which enable deter-
ting tool material used were found to be a decisive factor mination of the average interface temperature or the tem-
in the distribution of temperature fields and heat dissi- perature distribution curves along the tool–chip interface
pation [1,2]. In this regard, the application of tools [6,7]. This paper checks the fitting of temperature values
treated with advanced cutting tool coatings can lead to obtained experimentally by means of the natural thermo-
a substantial reduction of cutting temperatures and couple technique to the two different analytical models,
increase of the heat partition to the chip and, as a result, i.e. the power and linear relationships, which incorporate
to moderate tool wear [3,4]. This is due to the fact that basic thermal properties of the coupled materials under
both thermal and tribological outputs from the tool–chip actual interface conditions. It should be noted that the
exponential model was derived using the functional
analysis approach, originally proposed by Chao et al. [8]

Corresponding author. Tel.: +48-77-4006290; fax: +48-77- and developed later by Loewen and Shaw [9], and the
4006342. linear model was obtained based on the theory of simi-
E-mail address: grzesik@polo.po.opole.pl (W. Grzesik). larity [10].

0890-6955/$ - see front matter  2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0890-6955(03)00160-3
1312 W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317

In addition, this work considers heat partition in the The linear model involving thermophysical properties of
stationary tool and in the moving chip due to frictional the contacted work (W) and tool (T) materials is as fol-
heat at the tool–chip interface for two cases, namely lows:

冉 冊
orthogonal machining with carbide and multilayer
a0.625
⌰T ⫽ A ⫹B
W
coated tools. The novelty of this modeling approach (2)
results mainly from the mutual application of the Peclet l0.85
W lT
0.15

and Fourier numbers in one integrated model, which where aW is thermal diffusivity of the chip material, and
includes the changes in the thermodynamic properties of lW, lT are thermal conductivity of work and tool
the work (chip) and coating/substrate materials. (coating or substrate) materials.
It is reported [6,11] that the heat partition fraction for
a body with a stationary heat source (tool) decreases rap- 2.2. Determination of the heat partition coefficient
idly with an increase in the Peclet number, NPe. For com-
parison of this effect, a high Peclet number of NPe⬇ In this study, prediction of partition of the heat flux
5–20 is characteristic for conventional machining of which flows into the chip, i.e. for body with a moving
steel with a carbide tool, and a low value of NPe⬇0.5 is heat source, was based on the determination of the heat
typical for ultraprecision machining of aluminum with a partition coefficient (Rch), which defines the percentage
single-crystal diamond tool [12]. According to Reznikov of the heat entering the moving chip. It should be noted
[2], heat sources for which the Peclet criterion is higher that fraction (1⫺Rch) provides the percentage of the dis-
than 10 belongs to the class of high-speed heat sources. sipated energy going to the tool, i.e. the member that is
This means that the generated heat does not expand in stationary relative to the heat source. Basically, two dif-
its front parallel to the movement direction, but only ferent methods of calculation of Rch values are proposed.
below and beyond it. It is possible, with the increase Kato and Fujii [17] have determined this coefficient for
in the sliding velocity, to reduce substantially the heat conventional surface grinding based on the values of the
partition for the multilayer coated cutting tools tested. heat transmission coefficients (triple products used in Eq.
In the study carried out with steels and both uncoated (1)) for workpiece and grinding wheel materials. The
and coated carbide tools, the machining process was per- same approach was used in the ASM Surface Engineer-
formed at high values of the thermal number, which ing Handbook [18]. By analogy, the heat partition coef-
guarantees that the heat source moves faster than heat ficient for given thermal properties of work (W) and tool
can expand to the tool body [12,13]. (T) materials can be given by
1
Rch ⫽
1 ⫹ 冑(lrcp)T / (lrcp)W
(3)
2. Models for the prediction of interface
temperature and heat partition
On the other hand, Reznikov [2] has proposed calcu-
2.1. Prediction of the chip–tool interface temperature lation of the amount of heat flowing to the chip using
two dimensionless thermal numbers, namely the Peclet
As noted in Section 1 two models, power and linear, (Pe) and Fourier (Fo) numbers. The adequate expression
derived from the dimensional analysis and the theory of can be written as
similarity, are verified experimentally in the first part of 1
Rch ⫽
1 ⫹ (3 / 2)(lT / lW)冑PeTFoW
investigation. The models are given by Eqs. 1 and 2, (4)
respectively.
The Peclet number ascertains the velocity of the heat

冦 冧
n
source movement and for the uniform rectangular fric-

冪(lrc)
vch tional heat source localized on the tool–chip interface is
⌰T ⫽ CT ec (1)
W given by [5]
vchlc
Pe ⫽ (5)
aT
where ⌰T is mean tool face temperature [⌰], ec is spe-
cific cutting energy [FT2], vc is cutting speed [LT⫺1], h where vch is the sliding (chip) velocity, lc is the tool–
is undeformed chip thickness [L], and lWrWcp is heat chip contact length and aT denotes the diffusivity of the
transmission coefficient of work material [F2L⫺2T⫺1 tool material.
⌰⫺2], where lW, rW and cp are thermal conductivity, The Fourier number, which defines the duration of the
density and specific heat, respectively. contact (heat source influence), can be obtained as
Eq. (1) can be written in simplified form as: aWt aW
Fo ⫽ ⫽ (6)
⌰T ⫽ CTAn (1a) l2c lcvch
W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317 1313

where aW is the diffusivity of the work material. m / min for stainless steel, feed rate f = 0.16 mm/rev,
Combining Eqs. (4)–(6) leads to the final expression depth of cut ap = 2 mm.
for the heat partition coefficient: The measuring techniques were essentially similar to
those used in previous authors’ studies on cutting tool
1
Rch ⫽ coatings [5,14–16]. The cutting forces Fc and Ff were
1 ⫹ (3 / 2)(lT / lW)冑aW / aT
(7)
measured using a two-component strain-gauge dyna-
mometer fixed on the tool post of a lathe.
Fig. 1 illustrates the influence of the product PeTFoW on
For example, Fig. 2 shows values of the cutting forces
the heat partition for different values of the ratio lW / determined for varying cutting speed.
lT, ranging from 0.1 up to 2. It should be pointed out The thermal emf signals were recorded in the classical
based on Fig. 1 that the fraction of heat cumulated in tool–work thermocouple circuit and converted into equi-
the tool body (T) can be substantially reduced for large valent temperature values. In order to store the data in
values of the ratio of thermal conductivities of the sliding computer memory, on-line measurements were
partners and when the sliding system is itself able to accomplished using a multichannel data acquisition
provide a large thermal number or the maximum value software.
of the product PeTFoW. After cutting, the contact parts of the tool rake faces
On the other hand, the decrease of the ratio lW / lT were dimensioned with the PC-based optical image pro-
denotes that the majority of heat flux, i.e. qT = Rq, flows cessing system described in [15,16]. It included a CCD
to the tool with higher thermal conductivity. This situ-
camera, a high-resolution color display and a Corel
ation can occur when machining heat-resistant or high- Photo Paint v.7.0 graphical package. This system pro-
temperature alloys with carbide tools or plastics with vides a high-resolution and sharp image for visualization
diamond tools.
and, as a result, the tool–chip contact length was esti-

3. Experimental investigation

In this study, the cutting experiments were conducted


using two multilayers with an intermediate ceramic
layer: CVD—TiC/Al2O3/TiN—⌺8µm and TiC/Ti(C,N)/
Al203/TiN—⌺10µm coated flat-faced inserts consisting
of ISO-P20 cemented carbide substrate. In order to
compare experimental results, the substrate was also
examined as a tool material. The insert geometry was
ISO-TNMA 160408 with a clearance angle equal to 0°.
The cutting tests were carried out on a precision lathe
using a thin-walled tube as the workpiece. The cutting
was performed dry. The thickness of the tube wall was
equal to 2 mm and the outer diameter of the tube was
set at 80 mm. The work materials used in this study
were: AISI 1045 carbon steel (easy-to-machine material)
and AISI 304 austenitic stainless steel (difficult-to-
machine material). For all experiments, cutting para-
meters were varied over the ranges: cutting speed vc =
50–210 m / min for carbon steel and vc = 30–180 Fig. 2. Dependence of the cutting force on cutting speed.

Fig. 1. Influence of thermophysical properties of sliding partners—workpiece (W) and tool (T) on heat partition after [2].
1314 W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317

multilayer coated tools in the case of cutting AISI 304


stainless steel (courses 1b and 2b vs. 3b in Figs. 3 and 4).
For thermal analysis of the process, the thermophys-
ical properties of workpiece and coating materials occur-
ring at higher temperatures, including the thermal con-
ductivity (l) and the thermal diffusivity (a = l / cr), were
determined. For example, variations in a are partially
shown in Fig. 5(a). For the WC–Co carbides, a
decreases monotonically, but under higher contact tem-
perature exceeding 400 °C for both titanium-based coat-
ing materials and Al2O3-based ceramics, it changes
insignificantly. The distribution of the thermal properties
of the steels used depends strongly on their physical
properties. For austenitic stainless steel (Fig. 5b), both
the thermal diffusivity and conductivity increase slowly
with an increase in temperature. On the other hand, for
medium carbon steel, these properties decrease up to
720–730 °C and subsequently increase, as for stainless
steel. It should be pointed out that at the contact tempera-
tures occurring during the experiments, both l and a
Fig. 3. Dependence of the tool–chip contact length on cutting speed. changed rather linearly.

mated with high accuracy. Consequently, this processing


technique guarantees more accurate determination of the 4. Experimental and computation results
heat fluxes and the contact loads.
Fig. 3 shows changes in the tool–chip contact length The results of the interface temperature and the heat
resulting from the coatings used in the whole range of partition calculations according to the models described
cutting speeds applied. Moreover, Fig. 4 illustrates how in Sections 2.1 and 2.2 are presented successively in
the contact temperatures measured for the same cutting Figs. 6–10. In these figures, the experimental values of
speeds as in Fig. 3 influence the tool–chip contact Rch are presented for eight kinds of tool and work
behavior represented by the tool–chip contact length. It materials. It is evident from Fig. 6 that the temperature
is characteristic that substantial differences occur for values predicted by the linear model given by formula
(2) are consistent with the measured ones. Similar results
were obtained based on the power model defined by Eq.
(1). For example, for machining stainless steel with
three-layer coated tools under cutting parameters
applied, the empirical formulae are as follows:

冢 冣
0.43

冪(lrc)
vc h
⍜̄ ⫽ 27.03 ec in °C (8a)
W

⍜̄ ⫽ 10.342 ⫻ 108 冉 l
a
W
0.625
W
0.85 0.15
lT

⫺25.77 (8b)

⫻ 103 in °C
A slight disagreement between the experimental and cal-
culated values is observed. For these two models, calcu-
lation errors ranged from –3.0% to +1.54% and from –
1.03% to +0.65% for linear and polynomial approxi-
mations, respectively. In general, maximum calculation
error does not exceed 3.0% for all combinations of work
Fig. 4. Dependence of the tool–chip contact length on contact tem- and tool materials used.
perature. In contrast, Armarego et al. [20] reported that a com-
W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317 1315

Fig. 6. Relationship between measured and predicted values of the


tool–chip interface temperature for uncoated and multilayer coated
tools. Work material: AISI 304 stainless steel.

Fig. 7. Influence of the interface temperature on heat partition. 1—


Uncoated carbide P20, 2—TiC/Al2O3/TiN, 3—TiC/Ti(C,N)/Al2O3/
TiN, a—AISI 1045 carbon steel, b—AISI 304 stainless steel.

Fig. 5. Dependence of thermal diffusivity of contacted work (a) and


tool (b) materials on temperature rise [5,19].

prehensive predictive model for temperature and other


process outputs in ‘classical’ orthogonal cutting as a
function of major operation (technological) variables has
many limitations, and the predicted average tool–chip
interface temperature is on average within –10.7% and
51.7%, respectively, between the tool–work thermo-
couple and infrared methods. As a result, they concluded
that the quantitative correlation between the predicted
and measured average tool–chip interface temperature is
Fig. 8. Influence of the contact length on heat partition calculated
in very good agreement with the tool–work thermo-
using Eq. (7).
couple method.
1316 W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317

chip, being the member that is moving relative to the


frictional heat source, depending on the tool–chip inter-
face temperature, which varies from 430 to 930 °C. It
is evident from Fig. 7 that heat transfer to the chip
becomes more intensive at higher temperatures or equiv-
alently at higher cutting speeds. Moreover, the model
taking into account the Fourier number provides lower
values of the heat partition coefficient. The maximum
heat partition was revealed for machining carbon and
stainless steels with the three-layer coated tools. The
values of the R coefficient computed by Eq. (3) are equal
to 0.77 and 0.71, respectively. On the other hand, the
corresponding values of R determined by Eq. (7) are
Fig. 9. Influence of the thermal number on heat partition calculated equal to 0.68 and 0.62. Consequently, for uncoated car-
using Eq. (7).
bide tools, the partition of the heat flux varies from 0.51
to 0.42 depending on the combination of the steel grade
and the substrate/coating material. This implies that the
use of multilayer coated tools causes about 30% more
heat generated due to friction to be transferred into the
moving chip.
As shown in Fig. 8, the larger the length of the inter-
face contact, the higher the heat partition fraction for a
tool, which represents a body with a stationary heat
source. This means that less heat is carried away by the
chip, which is moving relative to the heat source. This
situation is mostly related to machining stainless steels,
for which the heat partition to the chip can be reduced
to 0.42 when using conventional carbide tools. One can
see based on Figs. 7 and 8 that the higher the velocity
Fig. 10. Influence of the specific friction energy on heat partition cal-
of the moving heat source, the closer the maximum value
culated using Eq. (7).
of temperature to the trailing edge. This experimental
evidence was also demonstrated by Komanduri and Hou
Also, the simulation results obtained by means of the [12] for sliding steel and bronze bodies.
finite difference method for the machining of mild steel Fig. 9 shows the variation of the heat partition coef-
with the feed rate of 0.604 mm/rev and cutting speed of ficient as a function of the Peclet number for different
22.86 m/min indicated that the predicted and measured workpiece and cutting tool materials. It can be reasoned
(classical Boothroyd’s data) values of the maximum based on Fig. 9 that the heat partition fraction qch =
temperature on the rake face were equal to 810 and 760 RqF increases distinctly with an increase in the thermal
°C, respectively [7]. number. In general, significant reduction of the heat par-
In this investigation, the heat partition coefficient was tition fraction for a body with a stationary heat source
analyzed in terms of fundamental contact characteristics (tool) results from a rapid increase in the Peclet number
such as the interface temperature, the tool–chip contact [12]. A similar effect was noted when considering the
length, the Peclet number and the specific friction influences of the cutting speed or cutting temperature.
energy. It should be noted that in Figs. 7–10 the heat As expected, the highest values of the heat partition for
partition expresses the percentage of friction energy the two sliding bodies occurred for coatings with an
going to the chip, i.e. R ⬅ Rch. The use of the interface intermediate ceramic layer and carbon steel used.
temperature as a process variable instead of the tradition- The comparison shown in Fig. 10 indicates that
ally used cutting speed is motivated by the fact that the substrate/coating–steel sliding systems, which provide
tool should not be employed close to or beyond the dif- minimum friction at the interface, are able to transfer
fusion and bonding limits of cutting tool materials. For more heat to the body with a moving heat source. It is
example, the diffusion limit for cobalt binding in tung- interesting to observe in Fig. 10 that this physical evi-
sten carbide tools is about 1100–1200 °C [7]. The rec- dence is mostly valid for machining stainless steel. It
ommended working temperatures for the range of coat- should also be noted based on Figs. 7 and 10 that the
ings mostly used in modern industry can be found in temperature rise for stainless steel is much higher than
[21]. that for carbon steel due to higher energy input (course
Fig. 7 shows the percentage of heat transferred to the 1b in Fig. 10) and relatively low thermal diffusivity
W. Grzesik, P. Nieslony / International Journal of Machine Tools & Manufacture 43 (2003) 1311–1317 1317

occurring at the interface temperatures lower than 600 [2] A.N. Reznikov, Thermophysical Aspects of Metal Cutting Pro-
°C (course 2 in Fig. 5). These temperature–energy cesses, Mashinostroenie, Moscow, 1981 (in Russian).
[3] W. Grzesik, Advanced Protective Coatings for Manufacturing
relationships are quite different from that observed in and Engineering, Hanser Gardner Publications, Cincinnati, 2003.
grinding operations performed on the same work [4] W. König, R. Fritsch, D. Kammermeier, Physically vapour
materials [17]. As reported by Varghese et al. [22] for deposited coatings on tools: performance and wear phenomena,
regular straight surface grinding of steels, energy par- Surf. Coat. Technol. 49 (1991) 316–324.
titions to the workpiece range from about 60% to 85% [5] W. Grzesik, P. Nieslony, M. Bartoszuk, Thermophysical-pro-
perty-based selection of coatings for dry machining of carbon and
with aluminum oxide, which, in general, is consist with stainless steels, Trans. N. Am. Manufact. Res. Inst. SME 30
the prediction results presented in this study. (2001) 343–350.
[6] R. Komanduri, Z.B. Hou, Tribology in metal cutting—some ther-
mal issues, Trans. ASME J. Tribol. 123 (2001) 799–815.
[7] L. Lazoglu, Y. Altinas, Prediction of tool and chip temperature
5. Conclusions in continuous metal cutting and milling, Int. J. Mach. Tools Man-
ufact. 42 (2002) 1011–1022.
1. Effects of the interface temperature (sliding speed), [8] R.T. Chao, K.J. Trigger, L.B. Zylstra, Thermophysical aspects of
metal cutting, Trans. ASME 74 (1952) 1039–1048.
length of the heat source, duration of the contact, and [9] E.G. Loewen, M.C. Shaw, On the analysis of cutting-tool tem-
thermophysical properties of two different materials perature, Trans. ASME 76 (1954) 217–231.
in sliding contact on the heat partition fraction were [10] S.S. Silin, Similarity Methods in Metal Cutting, Mashinostroenie,
investigated. Moscow, 1981 (in Russian).
2. It is possible to predict with reasonable accuracy the [11] R. Komanduri, Z.B. Hou, Analysis of heat partition and tempera-
ture distribution in sliding systems, Wear 251 (2001) 925–938.
interface temperature and heat partition based on the [12] R. Komanduri, Z.B. Hou, Thermal modeling of the metal cutting
contact characteristics and/or thermophysical proper- process—part II: temperature rise distribution due to frictional
ties of the work and substrate/coating materials. In heat source at the tool–chip interface, Int. J. Mech. Sci. 43 (2001)
particular, both power and linear models are accept- 57–88.
able for predicting the tool–chip interface tempera- [13] V.P. Astakhov, Metal Cutting Mechanics, CRC Press, Boca
Raton, 1999.
ture. [14] W. Grzesik, The role of coatings in controlling the cutting process
3. Incorporation of the Fourier number, which addition- when turning with coated indexable inserts, J. Mater. Process.
ally takes into account duration of the heat source Technol. 37 (1998) 133–143.
action, results in reduction of the heat partition frac- [15] W. Grzesik, The influence of thin hard coatings on frictional
tion for the body with a moving heat source. In gen- behaviour in the orthogonal cutting process, Tribol. Int. 33 (2000)
131–140.
eral, these discrepancies do not exceed 15%. [16] W. Grzesik, An integrated approach to evaluating the tribo-con-
4. Relatively strong influence of the contact character- tact for coated cutting inserts, Wear 240 (2000) 9–18.
istics, including the interface temperature, the tool– [17] T. Kato, H. Fujii, Energy partition in conventional surface grind-
chip contact length, the Peclet number and the specific ing, ASME Trans. J. Manufact. Sci. Eng. 121 (1999) 393–398.
friction energy, on the heat partition was observed in [18] S. Chandrasekar, T.N. Farris, et al., Thermal aspects of surface
finishing processes, in: Surface Engineering, ASM Handbook,
this study. vol. 5, ASM International, The Materials Information Society,
5. It was found that the use of multilayer coated tools USA, 1999, pp. 152–157.
causes about 30% more heat generated due to friction [19] F.P. Incropera, D.P. De Witt, Fundamentals of Heat and Mass
to be transferred into the moving chip. This is due to Transfer, third ed., John Wiley and Sons, New York, 1990.
the fact that the sliding system produces minimum [20] C.B. Aluwihare, E.J.A. Armarego, A.J.R. Smith, A predictive
model for temperature distribution in ‘classical’ orthogonal cut-
values of the specific frictional energy. ting, Trans. N. Am. Manufact. Res. Inst. SME 28 (2000) 131–
136.
[21] W. Grzesik, An investigation of the thermal effects in orthogonal
cutting associated with multilayer coatings, Ann. CIRP 50 (1)
References (2001) 53–56.
[22] V. Varghese, C. Guo, S. Malkin, G. Xiao, Energy partition for
[1] E.M. Trent, P.K. Wright, Metal Cutting, fourth ed., Butterworth- grinding of nodular cast iron with vitrified CBN wheels, Mach.
Heinemann, Boston, 2000. Sci. Technol. 4 (2) (2000) 197–208.

You might also like