Sunteți pe pagina 1din 11

Journal of Materials Processing Technology 211 (2011) 294304

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

An experimental and numerical study on the face milling of Ti6Al4V alloy:


Tool performance and surface integrity
Balkrishna Rao, Chinmaya R. Dandekar, Yung C. Shin
School of Mechanical Engineering, Purdue University, West Lafayette, IN 47907-2088, United States

a r t i c l e

i n f o

Article history:
Received 29 June 2010
Received in revised form 7 September 2010
Accepted 4 October 2010

Keywords:
Face milling
Titanium
Finite element analysis
Tool wear
Surface integrity

a b s t r a c t
This paper is concerned with the experimental and numerical study of face milling of Ti6Al4 V titanium
alloy. Machining is carried out by uncoated carbide cutters in the presence of an abundant supply of
coolant. Experimental analysis is conducted by focusing on the measurement of specic cutting energy,
surface integrity and tool performance. The experimental analysis is supplemented by simulations from a
3D nite element model (FEM) of face milling simulation where needed. A tool wear model parameterized
from FEM predictions of the toolchip interface temperature, contact stress and chip velocity is presented.
Tool wear patterns are described in terms of various cutting conditions and the inuence of tool wear
on surface integrity is investigated. Tool wear predictions based on the 3D FEM simulation show good
agreement with experimental tool wear measurements. The highest cutting speed realized for the cutting
tool material is 182.9 m/min (600 sfpm). Good surface integrity in terms of favorable residual stress and
surface nish is achieved under the machining conditions used with limited tool wear. Residual stresses
imparted to the machined surface are shown to be compressive.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Among the different alloys of titanium, Ti6Al4V is by far
the most popular one with its widespread use in the chemical,
surgical, ship building and aerospace industry. The primary reason for wide applications of this titanium alloy is its high
strength-to-weight ratio that can be maintained at elevated temperatures and excellent corrosion and fracture resistance. However,
Ti6Al4V is notorious for poor machinability due to its low thermal conductivity that causes high temperature on the tool face and
strong chemical afnity with most tool materials, thereby leading
to premature tool failure. Furthermore its inhomogeneous deformation by catastrophic shear makes the cutting force uctuate and
aggravates tool-wear and chatter. This poor machinability has limited cutting speed to less than 60 m/min in industrial practice as
described by Komanduri and Von-Turkovich (1981) and Chandler
(1989).
Therefore, over the years numerous research efforts have been
made to improve the machinability of Ti6Al4V by investigating tool-wear and related issues to assist in choosing suitable
machining conditions. Komanduri and Von-Turkovich (1981) studied the chip formation mechanism when machining Ti6Al4V and
reached a conclusion that prolonged contact between tool face and
chip underside during the upsetting stage under high temperature

Corresponding author. Tel.: +1 765 494 9775; fax: +1 765 494 0539.
E-mail address: shin@purdue.edu (Y.C. Shin).
0924-0136/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2010.10.002

conditions caused rapid tool wear due to Ti6Al4Vs chemical


afnity with most tool materials. This observation was later successfully utilized by Komanduri and Reed (1983) in designing a
new cutting tool holder that yields high clearance and negative rake
angles, which resulted in an improvement in tool-life. They also discovered the non-suitability of ceramics for machining Ti6Al4V. In
contrast, Narutaki and Murakoshi (1983) discussed the feasibility
of using natural diamond as a cutting tool material for machining
Ti6Al4V with an abundant coolant supply, and also described
the associated surface integrity. The effectiveness of using diamond and carbide cutters has also been described by Hartung and
Kramer (1982), who demonstrated the formation of a wear resistant titanium carbide reaction layer on these tool materials during
machining of Ti6Al4V. Later, Kramer and Chin (1993) used the
observations made in Hartung and Kramer (1982) to evaluate the
potential of various rare-earth metal compounds as tool materials
for machining titanium alloys.
All the work mentioned in the previous paragraph used low to
moderate cutting speeds with most of the tests performed under
orthogonal conditions. Industrial applications are, however, mostly
three-dimensional and require higher speeds whenever possible
to improve productivity. Due to the low thermal conductivity of
titanium alloys, controlling the heat during machining is vital in
successful machining of titanium. The interrupted nature of the
milling process does not allow build-up of heat energy generated
during cutting, which can be benecial from a tool-wear point of
view. Several researchers have worked along this line of thought.
Kitagawa et al. (1997) used a high cutting speed of 628 m/min in

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

end milling of Ti6Al4V with and without a coolant. This work


stressed the benets of improving tool-life by using intermittent
cutting along with a coolant. Kuljanic et al. (1998) studied the
possibility of end milling Ti6Al4V compressor blades with a polycrystalline diamond (PCD) cutter in the presence of coolant. The
authors arrived at an economical cutting speed of 110 m/min with
a tool life of 381 min. Recently, Barnett-Ritcey et al. (2001) investigated the use of a special coolant delivery system in high speed
milling of Ti6Al4V with polycrystalline diamond, coated carbide
and carbide cutters. Most of their tests were concerned with end
milling in the speed range of 152610 m/min. The authors presented the effect of coolant pressure and coolant aiming location
on tool-life. Ginting and Nouari (2006, 2007) investigated ball end
milling of Ti6242S and concluded that in the cutting speed range
of 60160 m/min carbide tooling is the most suitable. Other studies on ball end milling (Che Haron et al., 2007) and end-milling
(Lopez de Lacalle et al., 2000) with uncoated and coated carbide
tools concluded that in the cutting speed range of 40125 m/min
the coated carbide tool out-performed uncoated carbide in terms
of maximizing the tool life. In studies conducted on alternative
means of improving the tool life, Su et al. (2006) investigated the
performance of compressed cold nitrogen gas and oil mist cooling in end milling of Ti6Al4V with tool life almost 2.7 times
that of dry cutting and 1.9 times that of a coolant system with
only nitrogen-oil-mist at cutting speeds of 400 m/min. Hong et al.
(1998) demonstrated the capability of cryogenic cooling in milling
titanium showing promising results.
Surface integrity, which includes surface roughness, microstructure and residual stress after machining, is an important aspect of
successful titanium machining. Good surface integrity is especially
important in various engineering applications requiring high reliability and resistance to failure. It is generally understood that the
feed and tool nose radius play the most signicant role in surface
nish. In end-milling of titanium, Nurul Amin et al. (2007) showed
the dependence of surface roughness on the tool material. The surface roughness increased only marginally with an increase in tool
wear for uncoated carbides, while this effect was pronounced when
PCD inserts were used due to higher chatter observed at high cutting speeds. Contradictory results were reported by Elmagrabi et
al. (2008), who conducted slot milling experiments of titanium.
The authors concluded that for both uncoated and coated carbide tools the surface roughness was highly dependent on the feed
while tool wear did not seem to affect the surface roughness. It
has been shown that in milling of titanium compressive residual
stresses are generally observed as demonstrated by Mantle and
Aspinwall (2001) and Sun and Guo (2009). Mantle and Aspinwall
(2001) concluded that the resultant compressive residual stress
was dependent on ank wear and cutting speed. Increased tool
wear resulted in marginally higher compressive values of residual stress, while increased cutting speed reduced the compressive
residual stress. The contradictory effect of tool wear on surface
roughness and the lack of experimental data quantifying the residual stress as a function of the tool wear suggests further research is
needed to clarify this issue.
Compared to empirical or analytical methods, the advent of
computers has allowed researchers to study machining through
sophisticated numerical techniques. In numerical studies the primary focus has been on predicting the chip formation process and
cutting forces during turning of titanium. Umbrello (2008), Bker

295

et al. (2002), Bker (2006), zel et al. (2010a) and Calamaz et al.
(2008) all have conducted 2D orthogonal FEM studies on turning of Ti6Al4V, focusing on prediction of cutting force and chip
formation. Amongst them, Bker et al. (2002), zel et al. (2010a)
and Calamaz et al. (2008) focused on development of new material
models, while Hua and Shivpuri (2005) developed a tool wear prediction model. For milling, Ginting and Nouari (2006) conducted a
numerical study of dry milling in 2D approximation by qualitatively
comparing the predicted chip shape with experimental results
for varying cutting speeds and predicting the temperature using
the machining simulation software AdvantEdgeTM (Third Wave
Systems Inc, 2008). Recently, 3D FEM simulations have gained popularity in modeling machining processes such as turning, milling,
drilling, etc. Li and Shih (2006) simulated 3D turning of Ti6Al4V
using AdvantEdgeTM , focusing on the prediction of cutting forces,
temperature and the curling of the chip. Recently, Dandekar et al.
(2010) conducted 3D FEM turning simulations of Ti6Al4V to successfully predict the crater tool wear rate and the cutting forces. In
another study on turning, zel et al. (2010b) carried out 3D FEM
simulations to predict the cutting forces, chip morphology, temperatures and tool wear. Klocke et al. (2002) addressed some of
the challenges involved in selecting the right cutting parameters
through machining modeling for practical applications. The authors
suggested the utilization of the modeling outputs of stresses, relative velocities, temperature and strains to arrive at the possible
tool wear rate.
The majority of the work on 3D FEM of machining titanium is
focused primarily on turning and not on milling. Additionally, most
of the milling work on Ti6Al4V has focused on end-milling, while
little has been reported on other popular milling processes such as
face milling. Past studies have focused on tool wear and machinability issues with little reported on surface integrity resulting from
face milling of titanium alloys. The objective of this research is
to study face milling of Ti6Al4V by uncoated carbide cutters
via experiments and numerical modeling. The investigation is carried out in terms of measurement of the specic cutting energy
and mechanics, surface integrity and tool wear. The experimental
results are supplemented by numerical simulation results based
on a 3D FEM face milling model capable of predicting the specic
cutting energy, toolchip contact length, stress and temperature
distributions and are used in the parameterization of a tool wear
model. Tool wear is predicted based on 3D FEM face milling simulations, which has not been reported. Prediction of tool wear benets
the optimization of cutting parameters and also assists in designing
better tooling and cooling systems thereby improving the machinability of titanium.
2. Experimental equipment
Face milling tests were performed using a Mazak VQC-15/40
milling center with the maximum speed of 5000 rpm. These tests
were conducted on 50 mm 50 mm square blocks of Ti6Al4V
titanium alloy whose properties are listed in Table 1. The hardness
values listed in this table were obtained from a Mitutoyo hardness
tester (model ATK F1000). A standard face-milling cutter (Kennametal KDPR4SP430MB, lead of 30 , axial rake of 5 and radial
rake of 2 ) was used in the parametric study with uncoated carbide (K313) inserts. The cutting edge radius of the carbide inserts
was measured to be 24 m. It has been shown previously by Lopez

Table 1
Mechanical properties of Ti6Al4V.
No.

Work Material

Scale

Rockwell Hardness

Tensile Strength MPa

Yield Strength MPa

Ti6Al4V titanium

31.2 HRC

950 (36 HRC)a

880 (36 HRC)a

www.matweb.com.

296

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

CUTTER

EXIT
ENTRY

X
Y
WORKPIECE

MAZAK AXES

Fig. 1. Cutter and workpiece arrangement for machining of Ti6Al4V.

de Lacalle et al. (2000), Che Haron et al. (2007) and Dandekar et


al. (2010) that coated carbide tooling outperforms their uncoated
counterparts, but in this study due to restrictions on supplied tooling only uncoated carbide tooling was used. A schematic sketch
showing the cutter arrangement with respect to the workpiece is
shown in Fig. 1. All the tests were carried out using single-insert
y cutting. To observe associated tool wear behavior, fresh cutting
edges were used for all the cutting tests. A biostable heavy duty
water-soluble oil was used as the coolant with copious amounts
being directed towards the cutting edge. Unlike some of the previous research efforts (Barnett-Ritcey et al., 2001) where jet systems
were utilized to direct a pressurized coolant at the insert, the
present work directs a non-pressurized coolant stream at the cutting zone, as is available on most machines.
The cutting forces during the high speed face-milling tests were
measured with a Kistler 9257B dynamometer mounted under the

workpiece. The force signals were acquired by a LabVIEW based PC


data acquisition system at a sampling frequency of 1 kHz. The surface nish left on the machined workpieces was measured with
a portable SURTRONIC 3+ prolometer with a cutoff length of
0.8 mm.
The residual stresses imparted to the sub-surface layers were
measured by an X-ray diffraction device using copper radiation corresponding to a power supply of 20 mA and 40 kV. A Cu K radiation
source was used for a scanning range of 2 for each measurement
and was set between 139 and 145 to establish a window of angles
around the diffraction angle of 141.7 . The angle was varied for 6
equally spaced measurements in terms of sin2 for the measurement range of 045 . For peak tting a pseudo-Voigt function using
the software X-Fit was used. An average wavelength for the copper
radiation of  = 1.5417 A was used. A spot size of 0.1 mm diameter was used in all the measurements maintaining the irradiated
region as at as possible. Measurements were then conducted for
2 increment of 0.0014 at a speed of 0.025 /s. Machined workpieces were progressively etched to determine the depth proles
of the resultant residual stresses with a chemical etchant made of
distilled water (50%), nitric acid (40%) and hydrouric acid (10%).

3. 3D FEM machining simulations


The machining simulation software AdvantEdgeTM was used in
modeling of face milling using a single insert cutter. The 3D facemilling module with an indexable tool and focused cooling option
was used in the simulations, as this best describes the experimental
process (Fig. 2(a)). The updated-Lagrangian nite element method
along with continuous remeshing and adaptive meshing techniques was applied in the model. 4-node, 12 degree-of-freedom
tetrahedral nite elements were used to model the workpiece and
tool. A mesh convergence study was conducted rst to arrive at the
best possible mesh. In all simulations standard mode was used with
a minimum element edge length for the chip bulk and cutter edge as

Fig. 2. (a) Indexable face milling module in AdvantEdgeTM , (b) the focused cooling option: heat ux BC is applied to the region specied by user, (c) location of the jet center
on the rake face with a radius of 2 mm and (d) location of tool wear rate measurements on the ank face.

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

297

Table 2
JC model coefcients for Ti6Al4V.
Reference works

A (MPa)

B (MPa)

Khan et al. (2004)


Lee and Lin (1998)
zel and Karpat (2007) PSO-c optimization algorithm
zel and Karpat (2007) CPSO optimization algorithm

1080
782.7
883.9
987.8

1007
498.4
598.8
761.5

0.01304
0.028
0.0335
0.01516

0.6349
0.28
0.361
0.41433

0.77
1.0
1.041
1.516

0.0198 mm and 0.01 mm respectively. The radius of rened region


was set as 0.075 mm. A mesh renement factor of 5 was used and
the default factor of 5 was used for mesh coarsening. The boundary conditions were specied such that the top surfaces of the tool
are xed in vertical directions. The workpiece was constrained in
vertical and lateral directions on the bottom surface and the tool
rotates at the specied cutting speed.
The coolant was applied at the center of the rake face using the
focused coolant option allowable in AdvantEdgeTM as illustrated in
Fig. 2(b). The location of the coolant is illustrated in Fig. 2(c). A cutter
diameter of 10.16 cm (4 in.), the width of cut of 3 mm and a constant coefcient of friction of 0.65 corresponding to experimental
measurements were used in the simulations. The heat transfer coefcient was assumed to be 1500 W/m2 K corresponding to a heavy
duty water-soluble oil based cutting uid.

Fig. 3. Comparison between simulated and experimental tangential force for cutting
speed = 121.9 m/min, feed = 0.0508 mmpt and depth of cut = 0.762 mm.

3.1. Ti6Al4V material properties


In any FEM model an essential input is the accurate denition
of material properties. Under machining conditions, generally the
workpiece is subjected to extreme deformation involving high levels of strain, strain rates and rapid temperature rise. The workpiece
material is usually modeled by constitutive equations describing
the stressstrain response together with its dependence on strain
rate, temperature and work hardening. It was previously shown
that the default material model for titanium in AdvantEdgesTM
material library did not accurately represent the cutting forces
(Dandekar et al., 2010). Therefore, a user dened material model
was input into AdvantEdgeTM as shown in Eq. (1).

 = (A + Bnp ) 1 + C ln


o

 

 T T
r
Tm Tr

m 

(1)

where p and are strain and strain rate, o is the reference strain
rate (1/s) and n, m, A, B and C are constants determined through
experiments. Other material properties include: E = 114 GPa (the
modulus),  = 4428 kg/m3 (the density),  = 0.34 (the Poissons
ratio), = 9.6 106 / C (the coefcient of thermal expansion),
Tm = 1933 K (the melting temperature) and Tr = 296 K (the reference
temperature).
Table 2 summarizes the material constants obtained from the
literature for the JonhsonCook type constitutive model. Prior
to carrying out extensive simulations, a comparison was made
between the experimental tangential force and the simulated tangential force obtained by using the material model with various
materials constants as shown in Table 2. The nominal cutting
condition of speed = 121.9 m/min (400 sfpm), feed = 0.0508 mmpt
(0.002 ipt), and depth of cut = 0.762 mm (0.03 in.) was used for the
comparison.
As illustrated in Fig. 3 good results were obtained using the
material constants optimized by Khan et al. (2004) and zel and
Karpat (2007) using the PSO-c optimization algorithm. The nal
material model selected utilizes the material coefcients generated
by Khan et al. (2004) since the simulated tangential force is closest
to the experimental value.
Temperature-dependent thermal properties include specic
heat and thermal conductivity. Functions were tted to the available data and were included in the model.

For thermal conductivity (k, W/m K) (Mills, 2002):


If T 1260 K, k = 1.00 105 T2 0.0012T + 6.6519.
If T > 1260 K, k = 4.00 106 T2 0.0253T 6.269.
For specic heat (Cp , J/kg K) (Mills, 2002):
If T 1260 K, Cp = 0.21T + 483.37.
If T > 1260 K, Cp = 0.18T + 420.19.
4. Experimental matrix
In literature, use of positive rake angles for machining titanium
alloys has been suggested by earlier investigators (Barnett-Ritcey
et al., 2001). Accordingly preliminary milling tests were conducted using a positive rake angle cutter together with optimal
cutter entry and exit angles (Fig. 1) representing climb cutting. These tests allowed for machining to be carried out with
increased cutting speeds showing minimal tool wear. Therefore,
based on the results of the preliminary investigation, the nal
matrix for the uncoated carbide tool was designed and is shown
in Table 3. The nominal set of cutting condition selected are:
speed = 121.9 m/min (400 sfpm), feed = 0.0508 mmpt (0.002 ipt),
and depth of cut = 0.762 mm (0.03 in.). According to Barnett-Ritcey
et al. (2001), the nominal cutting speed selected happens to
be just above the threshold of 120 m/min (394 sfpm) for high
speed machining of Ti6Al4V. The highest cutting speed, highest feed and highest depth of cut employed for the carbide tool
Table 3
Experimental matrix cutter: KDPR-4-SP4-30MB, 4 in. diameter, 30 lead, 5 Axial
rake, 2 radial rake insert: SPG-422, 0.03125 in. nose radius.
Sr. no.
1
2
3
4
5
6
7
8
9

Tool material
Carbide
Carbide
Carbide
Carbide
Carbide
Carbide
Carbide
Carbide
Carbide

Speed m/min (sfpm)

Feed mmpt (in.)

Doc (mm)

76.2(250)
99.06(325)
110.6(362.8)
121.9(400)
152.4(500)
182.9(600)
121.9(400)
121.9(400)
121.9(400)

0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0889(0.0035)
0.127(0.005)
0.0508(0.002)

0.762
0.762
0.762
0.762
0.762
0.762
0.762
0.762
1.27

5.1. Mechanics of face milling of Ti6Al4V


The cutting forces measured during the cutting tests were transformed to the coordinate system rotating locally with the insert
as shown by Shin and Waters (1997) and Jensen and Shin (1999).
The specic cutting energy obtained from the tangential cutting
force is plotted in Fig. 4 as a function of cutting speed. Specic cutting energy was utilized as a metric to demonstrate its constant
behavior over the range of cutting speeds studied. Specic cutting
energy is a function of the cutting force and hence the comparison
between simulation and experimental results includes the cutting
force implicitly. With the carbide tool used, the tangential force
and hence the specic cutting energy remain nearly constant with
increasing cutting speed in the experimental range tested. Additional simulations were conducted at lower cutting speeds of 24.4,
36.6, 48.77 and 60.96 m/min to show the increase in the specic
cutting energy at lower cutting speeds. The simulation results capture this trend very well demonstrating the performance of the
material model input into the FEM.
This can be explained by observing the behavior of quantities
like stresses and temperatures in the primary and secondary deformation zones. Because it was not possible to measure quantities
such as cutting zone temperatures during the face milling operation, simulated results from FEA are used for analysis. For any FEA
software, its prediction accuracy is greatly dependent on the solution procedure employed and the underlying material constitutive
model. This accuracy was tested by comparing the predicted specic cutting energy and the thrust force with experimental data
as illustrated in Fig. 4. The error bars in all the gures stand for
the standard deviation of the measurements. The simulated specic cutting energy and the thrust force consistently over-predicts
the experimental results on an average by 6.2% and 14% respectively. Overall, the good agreement seen in this gure justies the
use of the FEA software for predicting temperature and stresses
generated in the cutting zone.
Fig. 5 presents the variation of the von Mises stresses and average temperatures, computed from FEA, with respect to cutting
speed in the primary shear zone. The error bars represent the standard deviation of the simulation results. The average temperatures
and stresses in the primary deformation zone show a nearly at
trend with increasing cutting speed. At the toolchip interface,
there is a slight increase in the average temperatures and the average stresses as shown in Fig. 6. However, both these quantities
remain nearly constant in the speed region between 121.9 m/min

800

1600

750

1400

700

1200

650

1000

600

800

550

600

500

400
Temperature

450

200

Mises Stress
400

0
50

75

100

125

150

175

200

Cutting Speed (m/min)


Fig. 5. Von Mises stress and temperature in the primary deformation zone at
feed = 0.0508 mmpt and depth of cut = 0.762 mm.

Average Temperature
Maximum Temperature
Mises Stress

1200

Temperature (deg C)

5. Results and discussion

Temperature (deg C)

are 182.9 m/min (600 sfpm), 0.127 mmpt (0.005 ipt) and 1.27 mm
(0.05 in.) respectively.

Von Mises Stress (MPa)

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

1000

1000

800

800
600
600
400
400
200

200
0

Von Mises Stress (MPa)

298

0
50

75

100

125

150

175

200

Cutting Speed (m/min)


Fig. 6. FEA predictions of stress and temperature in the secondary zone for carbide
tool at feed = 0.0508 mmpt and depth of cut = 0.762 mm.

and 182.9 m/min. The little change in the predicted values of the
toolchip interface stresses over a wide range of cutting speed is
also reected in the measured values of the coefcient of friction,
which vary between 0.62 and 0.68. Overall, the insignicant variation in stresses encountered in the deformation zones produces
cutting forces that remain unchanged over the range of cutting
speeds considered in this research.
Extremely short toolchip interface lengths have been observed
during the machining of Ti6Al4V. This is evident in Fig. 7, where
both the measured and simulated contact lengths are presented as
a function of cutting speed. The simulated chip length over-predicts
the experimental results on an average by 10.4% indicating a good

3000

70
60

2500

50

2000

40
1500

30

1000
500

Sim. Energy

Expt. Energy

20

Sim. Force

Expt. Force

10

0
20

45

70

95

120

145

170

195

Cutting Speed (m/min)


Fig. 4. Variation of specic cutting energy and thrust force with cutting speed at
feed = 0.0508 mmpt and depth of cut = 0.762 mm.

Tool chip contact length (mm)

80

Thrust Force (N)

Specific Cutting Energy


x 10^6 (J/m^3)

0.20
3500

Simulation
0.16

Experimental

0.12
0.08
0.04
0.00
50

75

100

125

150

175

200

Cutting Speed (m/min)


Fig. 7. Measured and predicted values of toolchip contact length for carbide tool
at feed = 0.0508 mmpt and depth of cut = 0.762 mm.

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

299

Fig. 8. Temperature ( C) and stress (MPa) distribution for two cutting speeds (Ti6Al4 V and carbide) at feed = 0.0508 mmpt and depth of cut = 0.762 mm.

5.2. Tool performance


Numerous face milling tests were conducted with the carbide
tool at different cutting speeds to study the progress of ank wear
with time. After each of these tests the ank wear value was determined by averaging wear-land widths measured at three separate
regions of the ank. The wear tests for each cutting speed were
continued until the failure of the cutting edge. The wear-time plots

so obtained are depicted in Fig. 9 with two sets of data graphed


for some of the cutting speeds to justify repeatability. The predominant wear mechanism observed in face milling of titanium was
chipping and aking of the carbide cutting tool as reported by Min
and Youzhen (1988), Machado and Wallbank (1990) and Jawaid et
al. (2000). Likewise, in this study it was observed during the wear
experiments that the cutting edge failed due to excessive chipping
around a ank wear value of 0.017 mm. The ank wear measured

0.02

Flank Wear (mm)

agreement between the two. High temperatures and high stresses


exist in these short toolchip contact regions with the maximum
temperature occurring in the close vicinity of the cutting edge as
shown in Fig. 8. This localized region, subjected to higher temperatures and higher stresses, near the tip causes damage to the tool
apex.
The FEM simulation results in Fig. 8 also show a comparison of
the temperature elds between two cases with low and high cutting
speeds. Compared with the low speed case of 76.2 m/min where the
high temperature region is limited to the tip of the tool, the case
for the high speed of 182.9 m/min shows a high temperature region
extending a little longer on the toolchip interface. The larger high
temperature regions on the rake face with increasing cutting speeds
correlate very well with the discolored patterns developed at the
tool rake face as will be shown in the tool wear section later in
Fig. 11.

0.015

76.2 m/min

0.01

99.1 m/min
110.6 m/min
0.005

121.9 m/min
152.4 m/min
182.9 m/min

0
0

20

40

60

80

Machining Time (min)


Fig. 9. Time history of ank wear for carbide tool at feed = 0.0508 mmpt and depth
of cut = 0.762 mm.

300

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

Fig. 10. Appearance of carbide cutting edge after the cutting distance of 55 mm at feed = 0.0508 mmpt and depth of cut = 0.762 mm.

was minimal but it provided to be a good indicator on the advent of


chipping. Hence this number has been chosen as the critical ank
wear value.
Since tool wear is the most important factor limiting tool life,
thus hampering employment of aggressive cutting conditions, the
wear mechanisms at play were studied through optical microscopy.
Fig. 10 presents two different views, for each cutting speed, of the

state of a fresh carbide cutting edge after engaging in the rst cut.
Table 4 lists the ank wear that was measured after this rst cut.
The ank wear values listed in Table 4 are very low. A fresh cutting
edge was used for each of the cases listed in Table 4.
Distinct patterns can be observed on the rake faces whereas
no trace of any crater wear can be seen on the rake face for the
range of cutting speeds presented in Fig. 10. Fig. 11 presents the

Table 4
Flank wear measured under the uncoated carbide cutting edge (rst cut, machined length = 55 mm).
Sr. no.

Speed m/min (sfpm)

Feed mmpt (ipt)

Doc (mm)

Flank wear mm (in.)

Expt. wear rate (mm/min)

Sim. wear rate (mm/min)

1
2
3
4
5
6
7
8
9

76.2(250)
99.06(325)
110.64(368)
121.9(400)
152.4(500)
182.9(600)
121.9(400)
121.9(400)
121.9(400)

0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0508(0.002)
0.0889(0.0035)
0.1270(0.005)
0.0508(0.002)

0.762
0.762
0.762
0.762
0.762
0.762
0.762
0.762
1.270

0.00665(0.00026)
0.00827(0.00032)
0.00903(0.00035)
0.01166(0.00046)
0.01433(0.00056)
0.01600(0.00063)
0.01367(0.00054)
0.01433(0.00056)
0.00968(0.00038)

0.000325
0.000604
0.000964
0.001620
0.005557
0.008472
0.008444
0.015582
0.001417

0.000293
0.000543
0.001055
0.002041
0.006259
0.007457
0.009771
0.017335
0.001338

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

301

wear is one of the major forms of tool wear in the machining of titanium alloys. The absence of crater wear is evidence of lower cutting
temperatures since crater wear occurs in that region of the rake
face where toolchip interface temperatures are high (Narutaki
and Murakoshi, 1983). There are two reasons for the lower temperatures on the rake face. First, the cooling action of the coolant
combats the high temperatures generated in the cutting zone. Second, the interrupted nature of face milling does not continuously
expose the rake face to higher pressures and temperatures as in
turning of this titanium alloy as explained by Komanduri and Reed
(1983) and Narutaki and Murakoshi (1983). The discolored regions
on the rake face coincide with the contact regions where the high
temperatures were developed. There is no discernable loss of the
cutting tool material, and this discolored region penetrates into the
cutting tool and becomes larger with cutting speed. These regions
are attributed to the diffusion of carbons from the carbide tool to
the chip under the high temperature (Min and Youzhen, 1988).
However with the progress of time, the width of the wear land
on the ank of carbide tool increases gradually. This can be seen in
Fig. 12 where images of the wear history have been captured for a
particular cutting speed. As seen in this gure, after the elapse of a
certain period of time, chipping of the cutting edge can be observed
besides increasing ank wear. And from this instance of time both
the cutting edge chipping and ank wear intensify gradually until
the cutting edge fails. The chipping of the cutting edge is initiated by
depletion of carbon from the tool substrate, which is carried away
by the moving chip (Ezugwu and Wang, 1997). The gradual depletion of carbon embrittles the cutting edge and this coupled with the
increasing forces due to ank-work interaction causes premature
failure at ank wear values as low as 0.018 mm.
5.2.1. Tool wear model
A tool wear model was parameterized to predict the tool wear
rate (dW/dt) by using the output from the FEM simulations along
the ank face of the normal stress ( t ), chip velocity (Vc ) and temperature () and inputting in Eq. (2). The nal reported value of
the simulated tool wear rate is the average of the 3 tool wear rate
readings taken along the lines indicated on Fig. 2(d). Although the
tool wear rate model implemented is an adhesion based model it
includes the effects of abrasion through the parameterization of
two constants C1 and C2 and the inclusion of a stress term. Furthermore Usui et al. (1984) arrived at the conclusion that at higher
temperatures the wear characteristic was the same no matter what
the type of wear; crater wear or ank wear. The model has been successfully implemented by Usui et al. (1984) for predicting crater and
ank wear rates.
dW
= t Vc C1 exp
dt

Fig. 11. Appearance of carbide rake face after multiple cuts with a feed
= 0.0508 mmpt and depth of cut = 0.762 mm.

state of the rake face after machining for a certain time beyond
the rst cut until tool failure. This gure also shows the total distance cut by each of these cutting edges before tool failure. Even
after multiple runs the cutting edge does not display any signs of
crater wear. The black patch seen on the toolchip contact area
for the speed of 121.9 m/min was caused by cutting edge chipping.
This is in contrast to observations by earlier researchers Hartung
and Kramer (1982) and Dearnley and Grearson (1986) that crater

C 
2

(2)

The drawback of the model is that the constants are dependent


on the combination of the tool material and the workpiece. To the
best of the authors knowledge there is no data available in the
literature for these constants for a combination of Ti6Al4V and
carbide. Therefore the constants were parameterized using experimental ank wear rate data for case no. 1, 4 and 6 corresponding
to cutting speeds of 76.2, 121.92 and 182.22 m/min respectively as
summarized in Table 4. The constants were calculated by plotting a
function of the wear rate (dW/dt), normal stress ( t ), chip velocity
(Vc ) against the inverse of the temperature (1/) as shown in Fig. 13.
The normal stress, chip velocity and the temperature readings were
obtained from the simulations for case no. 1, 4 and 6. Based on the
parameterization two constants were determined: C1 = 0.2955 and
C2 = 10098.
Using these values, tool wear rates were predicted for all the cutting conditions used in the experiments as summarized in Table 4.
For calibration of the model the cutting conditions referred to as

302

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

Fig. 12. Wear history of carbide tool at feed = 0.0508 mmpt, cutting speed of 99.1 m/min and depth of cut = 0.762 mm.

case no. 1, 4 and 6 were used, while for validation case no. 2, 3, 5, 7, 8
and 9 were used. The case no. corresponds to the serial no. in Table 4.
An expected result is the good agreement between the simulation
and experimental results of tool wear for the Vc of 99.06 m/min,
110.64 m/min and 152.4 m/min. This is apparent as the model was
calibrated for the cutting speed in the range of 76.2182.88 m/min.
Validation at Vc = 121.4 m/min, f = 0.0889 mmpt and 0.127 mmpt
and doc = 1.27 mm instills further condence in the model as model
calibrations were done using a f = 0.0508 mmpt and doc = 0.762 mm.
Additionally, care must be taken in interpreting the results as
the tool wear model was adjusted based on the experimental
data and used to predict the wear rate. On average, experimental measurements and the average predicted ank wear
rate differed only by 7%, thus showing a very good agreement
(Fig. 14).

5.3. Surface integrity


The impact of the face milling process on the integrity of the
machined surface was studied in terms of the residual stresses
and surface nish. Attention was focused on the tangential component of the residual stress which happens to be the major stress
component in the cutting direction. Besides, the tangential component allows for better averaging along a certain length of the
machined workpiece. Fig. 15 depicts a typical prole of the residual
stress measured as a function of depth below the machined surface.
As seen here and all the remaining stress data, these sub-surface
stresses are compressive in nature and their domain extends to
around 40 m below the machined surface.
The compressive nature of these stresses can be ascribed to the
intermittent nature of the face milling process in the presence of

Ln(dW/t*Vc*dt) (m^2/MN)

-8.5
182.88 m/min

Varying cutting
speed

-9.5
121.92 m/min

-10.5
y = -10098x - 1.219

-11.5
76.2 m/min

-12.5
0.0007

0.0008

0.0009

0.0010

0.0011

1/ (1/deg K)
Fig. 13. Determination of constants C1 and C2 in the tool wear rate model using
experimental ank wear data at feed = 0.0508 mmpt and depth of cut = 0.762 mm.

Fig. 14. Comparison of simulated and experimental measurements of tool wear rate
for face milling of titanium.

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

Surface finish (micron)

Stress (MPa)

-50
-100
-150
-200
-250
-300

303

10

12

14

16

18

20

76.2 m/min
99.1 m/min
121.9 m/min
152.4 m/min
182.9 m/min

0.8
0.6
0.4
0.2
0
0.005

22

0.01

Fig. 15. Typical residual stress prole for a carbide tool at speed = 121.9 m/min,
feed = 0.0508 mmpt and depth of cut = 0.762 mm.

coolant. Only the secondary cutting edge is in contact with the


machined surface during the cutting action of the face-milling cutter. The machined surface is also cooled by the coolant during other
times. Moreover, the ank wear amounts in all the cases were minimal, which also limited the input of heat ux into the sub-surface
layers of the milled workpiece. Thus, lesser heat energy is generated
at the secondary cutting edge during face milling and this leaves

0.015

0.02

0.025

Flank Wear (mm)

Depth Below Machined Surface (m)

Fig. 17. Surface nish as a function of ank wear for carbide tools at
feed = 0.0508 mmpt and depth of cut = 0.762 mm.

predominantly compressive residual stresses below the machined


surface due to reduced thermal impact (Brinksmeier et al., 1982).
Variation of the maximum residual stress value, for the carbide
tools, is presented in Fig. 16 as a function of cutting parameters,
where the maximum residual stresses change little with cutting
speed. As the cutting speed increases, the maximum compressive
residual stress decreases due to reduced time for heat conduction.
Increasing the feed leaves higher compressive residual stresses
below the milled surface as shown in Fig. 16. This is due to the
lower values of the frictional component of specic cutting pressure with increasing feed. The lower frictional pressure coupled
with the cooling action of coolant results in lower heat generation at higher feeds and this results in higher compressive stresses.
The trends depicted in Fig. 16 are similar to those obtained during
the high speed face milling of 7075-T6 aluminum (Rao and Shin,
2001). However, the stresses imparted to the titanium workpiece
are higher than their counterparts in 7075-T6 aluminum. This is
because of the higher yield strength of Ti6Al4V titanium when
compared to 7075-T6 aluminum. As for the nish imparted by the
face-milling process, the machined surfaces were very smooth with
most of the roughness values falling below 0.4 m.
5.3.1. Effect of tool wear on surface integrity
The variation in the surface nish and maximum residual stress
with increasing ank wear is presented in Figs. 17 and 18 respectively. For each cutting speed, there is no signicant change in
the nish left on the surface machined by the cutting edge which
is gradually wearing out. However, the maximum residual stress
imparted to the machined surface by a worn tool shows variation
with increasing ank wear. There is a reduction in the maximum
compressive stress value with the highest ank wear for each
speed. This is attributed to the increasing inux of heat energy into
the machined surface with increasing ank wear.
0

Maximum Stress (MPa)

-50

76.2 m/min
99.1 m/min

-100

110.6 m/min

-150
-200
-250
-300
-350
-400
0.005

0.007

0.009

0.011

0.013

0.015

0.017

0.019

Flank Wear (mm)


Fig. 16. Variation in maximum residual stress of Ti6Al4V.

Fig. 18. Maximum residual stress as a function of wear for carbide tools at
feed = 0.0508 mmpt and depth of cut = 0.762 mm.

304

B. Rao et al. / Journal of Materials Processing Technology 211 (2011) 294304

6. Conclusions
An experimental and numerical study of the face-milling of
Ti6Al4V titanium alloy with uncoated carbide tool materials was
undertaken. The following conclusions can be drawn from this
research:
1. There was very little variation in the specic cutting energy and
hence the cutting force and friction coefcient within the range
of cutting speeds studied in this research. In addition, very little
variation in the shear stresses and temperatures in the primary
shear zone has been observed from 3D FEM simulations.
2. In all the tests carried out in this work, the cutting edges for
the carbide tool materials suffered some damage due to the
high temperatures and pressures encountered on a very small
chiptool contact area. Tool failure occurred in the form of excessive chipping and notch wear.
3. A tool wear model for predicting the ank wear based on FEM
simulations was implemented. The predicted simulation results
compared well with experimental measurements.
4. In terms of the residual stress, at higher cutting speeds with carbide tooling, increasing feed increased the compressive residual
stresses left in the machined surface. On other hand, speed had
a negligible effect on residual stresses.
5. The surfaces machined by the carbide tool were very smooth
with most of the roughness values less than 0.4 m.
6. With increasing ank wear of the carbide tool, the nish
imparted by the worn tool to the machined surface remained
nearly constant while the maximum residual stress tends to
reduce.
References
Barnett-Ritcey, D.D., Hachmoller, R., Elbestawi, M.A., 2001. Milling of titanium alloy
using directed through spindle coolant. SME Technical Report. MR01-257, 18.
Bker, M., Rsler, J., Siemers, C., 2002. A nite element model of high speed metal
cutting with adiabatic shearing. Computers and Structures 80, 495513.
Bker, M., 2006. Finite element simulation of high-speed cutting forces. Journal of
Materials Processing Technology 176, 117126.
Brinksmeier, E., Cammett, J.T., Konig, W., Leskovar, P., Peters, J., Tonshoff, H.K., 1982.
Residual stresses measurement and causes in machining processes. Annals of
the CIRP 31, 491510.
Calamaz, M., Coupard, D., Girot, F., 2008. A new material model for 2D numerical simulation of serrated chip formation when machining titanium alloy Ti6Al4V.
International Journal of Machine Tools and Manufacture 48 (34), 275288.
Chandler, H.W., 1989. Machining of Reactive Metals. ASM Handbook Machining,
vol. 10, pp. 844857.
Che Haron, C.H., Ginting, A., Ashad, H., 2007. Performance of alloyed uncoated and
CVD-coated carbide tools in dry milling of titanium alloy Ti6242S. Journal of
Materials Processing Technology 185, 7782.
Dandekar, C.R., Shin, Y.C., Barnes, J., 2010. Machinability improvement of titanium alloy (Ti6Al4V) via LAM and hybrid machining. International Journal
of Machine Tools and Manufacture 50 (2), 174182.
Dearnley, P.A., Grearson, A.N., 1986. Evaluation of principal wear mechanisms of
cemented carbides and ceramics used for machining titanium alloy IMI318.
Materials Science and Technology 2, 4758.
Elmagrabi, N., Che Hassan, C.H., Jaharah, A.G., Shuaeib, F.M., 2008. High speed milling
of Ti6Al4V using coated carbide tools. European Journal of Scientic Research
22 (2), 153162.
Ezugwu, E.O., Wang, Z.M., 1997. Titanium alloys and their machinability a review.
Journal of Materials Processing Technology 68, 262274.
Ginting, A., Nouari, M., 2006. Experimental and numerical studies on the performance of alloyed carbide tool in dry milling of aerospace material. International
Journal of Machine Tools and Manufacture 46, 758768.
Ginting, A., Nouari, M., 2007. Optimal cutting conditions when dry end milling the
aeroengine material Ti6242S. Journal of Materials Processing Technology 184,
319324.
Hartung, P.D., Kramer, B.M., 1982. Tool wear in titanium machining. Annals of the
CIRP 31 (1), 7580.

Hong, S.Y., Qu, X., Lee, A., 1998. Economical cryogenic milling for environmentally
safe manufacturing. In: Proceedings of the NAMRI/SME, May, Atlanta, USA.
Hua, J., Shivpuri, R., 2005. A cobalt diffusion based model for predicting crater wear
of carbide tools in machining titanium alloys. Transactions of the ASME Journal
of Engineering Materials and Technology 127 (1), 136144.
Jawaid, A., Sharif, S., Koksal, S., 2000. Evaluation of wear mechanisms of coated
carbide tools when face milling titanium alloy. Journal of Materials Processing
Technology 99, 266274.
Jensen, S.A., Shin, Y.C., 1999. Stability analysis in facemilling operation: part 1. Theory
for stability lobe prediction. Transactions of the ASME, Journal of Manufacturing
Science and Engineering 21 (4), 600605.
Khan, A.S., Suh, Y.S., Kazmi, R., 2004. Quasi-static and dynamic loading responses
and constitutive modeling of titanium alloys. International Journal of Plasticity
20, 22332248.
Kitagawa, T., Kubo, A., Maekawa, K., 1997. Temperature and wear of cutting tools in
high speed machining of Inconel 718 and Ti6Al6V2Sn. Wear 202, 142148.
Klocke, F., Markworth, L., Messner, G., 2002. Modeling of Ti6Al4V machining operations. In: Proceedings of the 5th CIRP International Workshop on Modeling of
Machining Operations, Lexington, KY, pp. 6370.
Komanduri, R., Von-Turkovich, B.F., 1981. New observations on the mechanism of
chip formation when machining titanium alloys. Wear 69, 179188.
Komanduri, R., Reed, W.R., 1983. Evaluation of carbide grades and a new cutting
geometry for machining titanium alloys. Wear 92, 113123.
Kramer, B.M., Chin, V.S., 1993. Theoretical consideration of rare earth metal compounds as tool materials for titanium machining. Annals of the CIRP 42,
111114.
Kuljanic, E., Fioretti, M., Miani, F., 1998. Milling titanium compressor blades with
PCD cutter. Annals of the CIRP 47 (1), 6164.
Lee, W.-S., Lin, C.-F., 1998. Plastic deformation and fracture behavior of Ti6Al4V
alloy loaded with high strain rate under various temperatures. Materials Science
and Engineering A 241, 4859.
Li, R., Shih, A.J., 2006. Finite element modeling of 3D turning of titanium. International Journal of Advanced Manufacturing Technology 29, 253261.
Lopez de Lacalle, L.N., Perez, J., Llorente, J.I., Sanchez, J.A., 2000. Advanced cutting
conditions for the milling of aeronautical alloys. Journal of Materials Processing
Technology 100, 111.
Machado, A.R., Wallbank, J., 1990. Machining of titanium and its alloysa review.
Proceedings of the Institution of Mechanical Engineers, Part B, Management and
Engineering Manufacture 204 (1), 5360.
Mantle, A.L., Aspinwall, D.K., 2001. Surface integrity of a high speed milled gamma
titanium aluminide. Journal of Materials Processing Technology 118, 143150.
Mills, C.K., 2002. Recommended Values of Thermophysical Properties for Selected
Commercial Alloys. Woodhead Publication, Abington.
Min, W., Youzhen, Z., 1988. Diffusion wear in milling titanium alloys. Materials
Science and Technology. 4, 548553.
Narutaki, N., Murakoshi, A., 1983. Study on machining of titanium alloys. Annals of
the CIRP 32 (1), 6569.
Nurul Amin, A.K.M., Ismail, A.F., Nor Khairusshima, M.K., 2007. Effectiveness of
uncoated WCCo and PCD inserts in end milling of titanium alloyTi6Al4V.
Journal of Materials Processing and Technology 192193, 147158.
zel, T., Karpat, Y., 2007. Identication of constitutive material model parameters
for high strain rate metal cutting conditions using evolutionary computational
algorithms. Materials and Manufacturing Processes 22, 659667.
zel, T., Sima, M., Srivastava, A., 2010a. Finite element simulation of high speed
machining Ti6Al4V alloy using modied material models. Transactions of
NAMRI/SME 38, 4956.
zel, T., Sima, M., Srivastava, A.K., Kaftanoglu, B., 2010b. Investigations on the effects
of multi-layered coated inserts in machining Ti6Al4V alloy with experiments
and nite element simulations. Annals of the CIRP-Manufacturing Technology
59 (1), 7782.
Rao, B., Shin, Y.C., 2001. Analysis on high speed face-milling of 7075-T6 aluminum
using carbide and diamond cutters. International Journal of Machine Tools And
Manufacture 41 (12), 17631781.
Shin, Y.C., Waters, A.J., 1997. A new procedure to determine instantaneous cutting force coefcients for machining force prediction. International Journal of
Machine Tools and Manufacture 37 (9), 13371351.
Su, Y., He, N., Li, L., Li, X.L., 2006. An experimental investigation of effects of cooling/lubrication conditions on tool wear in high-speed end milling of Ti6Al4V.
Wear 261, 760766.
Sun, J., Guo, Y.B., 2009. A comprehensive experimental study on surface integrity
by end milling Ti6Al4V. Journal of Materials Processing Technology 209,
40364042.
Third Wave Systems Inc., 2008. AdvantEdgeTM FEM 5.1 Users Manual. Third Wave
Systems Inc, Minneapolis.
Umbrello, D., 2008. Finite element simulation of conventional and high speed
machining of Ti6Al4V alloy. Journal of Materials Processing Technology 196,
7987.
Usui, E., Shirakashi, T., Kitagawa, T., 1984. Analytical prediction of cutting tool wear.
Wear 100, 129151.

S-ar putea să vă placă și