Você está na página 1de 7

Journal of Materials Processing Technology 209 (2009) 40364042

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

A comprehensive experimental study on surface integrity by


end milling Ti6Al4V
J. Sun, Y.B. Guo
Department of Mechanical Engineering, The University of Alabama, Tuscaloosa, AL 35487, USA

a r t i c l e

i n f o

Article history:
Received 21 April 2008
Received in revised form
13 September 2008
Accepted 18 September 2008
Keywords:
Titanium Ti6Al4V
Surface integrity
Surface roughness
Residual stress
Milling

a b s t r a c t
End milling titanium Ti6Al4V has wide applications in aerospace, biomedical, and chemical industries.
However, milling induced surface integrity has received little attention. In this study, a series of end
milling experiments were conducted to comprehensively characterize surface integrity at various milling
conditions. The experimental results have shown that the milled surface shows the anisotropic nature
with the range of surface roughness values from 0.6 to 1.0 m. Surface roughness value increases with
feed and radial depth-of-cut (DoC), but has much less variation in the cutting speed range. Compressive
residual normal stress occurs in both cutting and feed directions, while the inuences of cutting speed and
feed on residual stress trend are quite different. The microstructure analysis shows that phase becomes
much smaller and severely deformed in the near surface with the cutting speed, but phase transformation
was absent for the milling conditions. The milled surface microhardness is about 7090% higher than the
bulk material in the subsurface.
2008 Elsevier B.V. All rights reserved.

1. Introduction

1.1. Surface roughness

Titanium alloys are widely used in aerospace, biomedical, and


chemical industries primarily due to the exceptional strength
to weight ratio (Ezugwu, 2005), high temperature performance
(Zoya and Krishnamurthy, 2000) and corrosion resistance (Copper,
2006). For example, almost all titanium monolithic components
in aerospace industry are manufactured by milling (Sun and Guo,
2008). However, titanium alloys are typically difcult-to-cut materials due to the high strength at elevated temperatures (Kahles et al.,
1985), low modulus, low thermal conductivity (Hong et al., 1993)
and high chemical reactivity (Su et al., 2006). The machined titanium components such as compressors of aircraft engines require
the greatest reliability which is determined by the process induced
surface integrity. However, surface integrity of milled titanium
components easily deteriorates due to the poor machinability of
titanium alloys and cyclic chip loading during milling. Milling
induced surface integrity, including anisotropic surface roughness,
residual stress, surface microstructure alterations and microhardness, has received little attention. The state-of-the-art knowledge
on surface integrity of machined titanium components is assessed
as follows.

Machined surface roughness depends on several factors, such


as cutting speed, tool wear, feed, tool materials and geometry, etc.
(Amin et al., 2007). Zoya and Krishnamurthy (2000) reported that
surface roughness value decreased with the increase of cutting
speed in the range of 150185 m/min, but increased in the range
of 185350 m/min when turning + phase stabilized titanium
alloys. Several researchers found that surface roughness values
became larger at high cutting speeds in turning Ti6Al4V (Ribeiro
et al., 2003) and in end milling Ti6Al4V using WC-Co and PCD
insert (Amin et al., 2007). As for the inuence of tool wear on surface roughness, Lopez de lacalle et al. (2000) found that with the
increase of cutting speed, surface roughness value rst increased
then decreased with the tool wear progression in milling using
hard solid mills. Che-Haron and Jawaid (2005) observed the similar
phenomena that the surface tends to become smoother toward the
end of tool life when turning Ti6Al4V. On the contrary, Canteroa
et al. (2005) reported that surface roughness value increased near
tool failure (in terms of ank wear) in drilling Ti6Al4V. On the
side of depth-of-cut (DoC) effect, Lopez de lacalle et al. (2000)
found that an increase of radial DoC caused an increase of surface
roughness value, while axial DoC has little inuence. Hughes et al.
(2004) showed that an increase in feed rate resulted in a larger
surface roughness value due to more feed marks. Ezugwu et al.
(2007) observed that surface nish was not affected by coolant
pressures.

Corresponding author. Tel.: +1 205 348 2615; fax: +1 205 348 6419.
E-mail address: yguo@eng.ua.edu (Y.B. Guo).
0924-0136/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2008.09.022

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

4037

Table 2
Residual stress measurement parameters.

Table 1
Milling conditions.
Experiment
No.

Cutting speed
V (m/min)

Feed f
(mm/tooth)

Radial DoC ae
(mm)

Axial DoC ap
(mm)

1
2
3
4
5
6
7
8
9
10
11
12
13

50
65
80
95
110
65
65
65
65
80
80
80
80

0.08
0.08
0.08
0.08
0.08
0.06
0.1
0.12
0.14
0.08
0.08
0.08
0.08

4
4
4
4
4
4
4
4
4
2
3
5
6

1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5

Sin2

Measurement method
Radiation
Spot size (mm)
Voltage (kV)
Amperage (mA)
( )
2 ( )

Co K
0.8
40
35
10, 20, 30, 40, 0, 10, 20, 30, 40
47.45

close to drilled surface. On the other hand, Ezugwu et al. (2007)


reported a softened surface by turning. Warren and Guo (2006)
have shown that the apparent paradox of surface softening was
caused by the edge effect on a measured surface during micro
indentation.

1.2. Microstructure changes

1.4. Residual stresses

Hughes et al. (2004) observed that phase deformation consisted of deformed boundaries and elongation grains in turning
Ti6Al4V. Che-Haron and Jawaid (2005) related the phase
deformed layer with tool wear conditions. The deformed layer was
very thin using a sharp tool while it is thick when the tool is nearly
worn out. Reissig et al. (2004) also observed subsurface deformation in drilling Ti6Al4V. Ezugwu et al. (2007) pointed out that in
ne cutting conditions, the shearing forces cannot cause severe
plastic deformation in the subsurface regardless of conventional or
high pressure cooling condition.
Besides phase deformation, microstructural changes such as the
absence of phase close to the hole surface have been shown in
drilling Ti6Al4V (Canteroa et al., 2005). The SEM-EDS analysis
showed that there exists an absence of phase close to the hole
surface. Similar results were also reported by Li et al. (2007). The
phase transformation from phase to phase was induced by high
temperatures resulted from the severe plastic deformations.

An early study by Zlatin and Field (1973) showed that an


abusive milling produced tensile residual stress in the machined
layer, while a gentle milling produced compressive residual stress.
Norihiko et al. (1983) reported that surface residual stress increased
with the increase of cutting speed. The magnitudes of residual
stresses by dry cutting Ti6Al4V were also larger than those by
wet cutting. Yang et al. (2002) found that residual stress on the
ground Ti6Al4V surfaces have a larger scatter than that of the
face turned ones in a statistical sense.
The above literature review shows that the bulk study of
machining Ti6Al4V is limited to either turning or individual
surface integrity factors in milling. A comprehensive study on
milling induced surface integrity including anisotropic surface
roughness, residual stress, microstructure alterations, microhardness and the relationship between these factors is signicantly
lacking. In addition, there are many inconsistent and even contradictory results in surface integrity. The objective of this study is to
ll in the knowledge gap and solve the impressing issues in milling
Ti6Al4V.

1.3. Subsurface microhardness


Recent research (Che-Haron and Jawaid, 2005; Hughes et al.,
2004) have shown that a work hardened surface layer was up to
100 m and associated with higher hardness than the bulk material, while the subsurface hardness at about 200 m was less than
the bulk material in the turned Ti6Al4V surface. Canteroa et al.
(2005) also reported a 30% hardness increase in the deformed layer

2. Experimental setup and procedures


The work material is Ti6Al4V alloy (ASTM B265 grade 5 titanium). A series of down milling tests were performed on a CNC
milling center. The 12.7 mm diameter cutting tool is a 4-ute solid
carbide end mill with TiAlN coating. The down milling tests were

Fig. 1. Residual stress measurement setup and directions.

4038

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

Fig. 2. Schematic of diffraction planes.

carried out with water-soluble cutting uid. The detail milling conditions are listed in Table 1. Five levels of cutting speed, feed, and
radial DoC were selected, while the axial DoC was kept a constant
value. A sharp cutting edge was used at milling condition to ensure
measurement repeatability and fair comparison.

A series of surface integrity characterizations were conducted.


Surface topography and roughness were measured by using an optical microscopy and a stylus proler. SEM images of the subsurface
microstructure were also taken. Surface and subsurface microhardness (HK) was measured using a microhardness indenter with the
load of 25 g.
XRD-based residual stress measurement was used to measure
surface residual stress in the cutting, feed and shearing directions
in Fig. 1. The residual stress measurement parameters are listed
in Table 2. The XRD method is based on the fact that strain in
a crystalline material is induced as a consequence of mechanical
deformation, phase transformation, thermal expansion, etc. The
strain causes a change in the atomic spacing within the crystal
structure compared with the stress-free condition (Snoha, 1996;
Hauk, 1997). The change of inter-atomic spacing d causes a shift in
the diffracted X-ray peak position. By resolving the angular peak
shift and applying the Bragg law to quantify the d-spacing, residual
stress on the machined surface can be calculated using the theory
of linear elasticity. Plane stress condition on the measured surface
was assumed, i.e., only residual normal stress 11 and 22 and
shear stress 12 exist but 33 perpendicular to the surface is zero.
The common sin2 method was used in the determination of
residual stress. The sin2 method requires a number of XRD measurement made at different tilts, Fig. 2 (Fitzpatrick et al., 2002). The
angular position of the diffracted peak was determined by the least-

Fig. 3. The effect of radial DoC (ae ) on surface topography (V, 80 m/min; f, 0.08 mm/tooth; ap = 1.5 mm).

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

squares parabolic curve tting method to calculate d-spacing from


the Bragg relation. A plot of d-spacing vs. sin2 was constructed.
The slope of a least-square line tted to the experimental data multiplied by the X-ray elastic constant (E/(1 + )) is proportional to the
stress on the plane of the surface. A linear d-spacing vs. sin2 plot
indicates that the strain distribution is homogeneous within the Xray irradiated volume and that the assumption of plane stress state
is valid (Snoha, 1996; Fitzpatrick et al., 2002).
3. Surface integrity characterization
Machining induced surface integrity can be generally described
by its topological, mechanical, metallurgical and chemical states.
In this study, surface integrity of the milled surfaces was systematically characterized by surface roughness, residual stress,
microstructure and microhardness. The coherent characterizations
provide a physical basis for the understanding of machined component performance such as fatigue life and stress corrosion in
service.

4039

in both directions is within 0.61.0 m, while roughness value is


slightly larger in cutting direction than that in feed direction at relatively high speeds. The feed effect on surface roughness value in
Fig. 5b shows that it almost linearly increases with the increased
feed in both directions. The variation range of surface roughness
value in both is 0.61.2 m and, therefore, is equivalent. As for the
effect of radial DoC in Fig. 5c, surface roughness value increases
with the increased radial DoC in trend which is consistent with the
observed phenomenon in Fig. 3. Surface becomes rougher in cutting
direction at low radial DoCs, while it is smoother at relatively high
radial DoC. However, surface roughness value is also in the range of
0.61.2 m in both directions, which is equivalent to Fig. 5a and b.
The sensitivity of surface roughness value to process parameters
shows that the feed has the most signicant inuence, followed by
radial DoC, while cutting speed has the least inuence. Although a
certain variation of surface roughness value exists for the milling
conditions under consideration, it is generally in the range of
0.61.0 m.
3.2. Residual stress

3.1. Surface roughness


The inuence of radial DoC on surface topography is shown in
Fig. 3. It can be seen that the surface becomes rough when the radial
DoC increases due to the increased overlap between cutting paths.
It can also be induced that the increase of feed also results in more
surface roughness value. Fig. 4 shows the simulated surface pattern and the optical image of the experimental surface pattern. The
simulated pattern based on process parameters and tool diameter
can characterize the basic features of the measured one. It should
be pointed out that the simulated surface pattern is only based on
tool path kinametics, while the surface material deformation was
not incorporated. Therefore, the difference between the simulated
pattern and the experimental one is expected. However, the simulated pattern still shows the basic characteristics of the milled
surface topography.
As shown in Fig. 3, surface topography in the cutting and
feed directions shows the anisotropic nature, which means surface roughness values will be different in the different directions.
The measured arithmetic mean Ra in cutting and feed directions
is shown in Fig. 5. Surface roughness value in cutting direction,
Fig. 5a, initially increases with the cutting speed (up to 80 m/min)
and then decreases with the further increased cutting speed. In the
feed direction, surface roughness value only slightly decreases with
the increased cutting speed. The range of surface roughness value

The effect of cutting speed on residual stress at the surface in


cutting and feed directions is shown in Fig. 6. For repeatability,
three measurements at different locations for each surface were
made and the results were averaged to give the residual stress
proles. In feed direction, compressive residual normal stress 11
increases with cutting speed, reaches to the maximum 350 MPa
around 80 m/min, then reduces with the increased cutting speed
in Fig. 6a. In cutting direction, residual normal stress 22 shows a
similar trend to 11 but with less magnitude of about 100 MPa at
lower cutting speeds. Compared with the normal residual stresses,
shear stress 12 shifts from tension to slight compression with the
increased cutting speed. In addition, the magnitudes (100 MPa to
100 MPa) of shear residual stress are much smaller than those of
normal residual stress.
The underlying mechanism for the residual stress variation is
due to the combined effects of mechanical and thermal deformations. The dominant mechanical deformation at all cutting
speeds tends to produce compressive residual stress. At low cutting speeds, the degree of mechanical deformation is not as large as
that at high cutting speeds. The magnitude of compressive residual stress increases with cutting speeds. However, when cutting
speed increases over 80 m/min, the increased cutting temperature
induces thermal deformation which tends to produce tensile residual stress and, therefore, reduce the magnitude of compressive

Fig. 4. The inuence of tool path on surface pattern (V, 80 m/min; f, 0.08 mm/tooth; ae = 3 mm; ap = 1.5 mm). (a) Cutting speed effect (f, 0.08 mm/tooth; ae = 4 mm; ap = 1.5 mm).
(b) Feed effect (V, 65 m/min; ae = 4 mm; ap = 1.5 mm). (c) Radial DoC effect (V, 80 m/min; f, 0.08 mm/tooth; ap = 1.5 mm).

4040

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

residual stress. The highly nonlinear coupling of mechanical and


thermal deformations determines the characteristics of the residual
stress proles.
Fig. 6b shows that compressive residual normal stress generally
decreases with the feed. The variation of residual stress at the feed
of 0.08 mm/tooth does not affect the curve trends. The shear residual stress becomes more compressive when the feed increases. The
magnitude (0 to 100 MPa) of residual shear stress is much smaller
than the residual normal stresses, especially at relatively low feeds.

Fig. 6. The inuence of milling parameters on surface residual stress (V, 65 m/min;
ae = 4 mm; ap = 1.5 mm).

The variations of the residual stresses are expected due to the


non-uniform surface deformations as shown in Figs. 3 and 4. The
surface asperities or roughness indicate that the material deformations at these locations are different and, therefore, result in
a certain variation of residual stress. Since the X-ray spot on the
milled surface covers a certain area including different asperities,
the measured residual stress is an average residual stress of the
measured area. In addition, a number of error sources, such as the
method of curve tting to test data, measurement repeatability,
and various sources of uncertainty, contribute to the variation of
residual stress.
3.3. Subsurface microstructure

Fig. 5. The inuence of milling parameters on surface roughness. (a) Cutting speed
inuence (f, 0.08 mm/tooth; ae = 4 mm, ap = 1.5 mm). (b) Feed inuence.

The cross-section views of subsurface microstructure are shown


in Fig. 7. At low cutting speeds (50 and 65 m/min), phase change
(deformation or volume change) in the near surface cannot be
observed. When the speed increases to 80 m/min, the deformation of phase in the near surface is clear. At 95 m/min, the
volume of phase becomes very small and almost disappears in
the near surface. Similarly, the volume of phase also seems to be
smaller and deformation looks larger at the speed of 110 m/min
than those at low speeds of 50, 65 and 80 m/min. The varia-

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

4041

Fig. 7. Cross-section view of subsurface microstructure (f, 0.08 mm/tooth; ae = 4 mm; ap = 1.5 mm).

tion of near surface microstructure is caused by the combined


effects of mechanical and thermal loads in milling. The variation trend is qualitatively consistent with that of residual stress
in Fig. 6a. However, a quantitative relationship between the near
surface microstructure and residual stress needs a further study in
the future.
Since the surfaces were milled with the sharp cutting tool at
relatively low cutting speeds and well-lubricated conditions, cutting temperatures are too low to induce any phase transformations
such as a white layer on the machined surfaces. It implies that no
thermal damage occurred at the milling conditions. Surface deformation induced strain hardening should be the controlling factor
for the measured surface hardness in the next section.
3.4. Microhardness
Fig. 8 shows the surface and subsurface Knoop microhardness
distributions. Each microhardness prole represents the averaged four sets of microhardness data. The microhardness proles
are characterized by the higher surface hardness (13501500 HK)

Fig. 8. Microhardness distribution and variation in the subsurface (f, 0.08 mm/tooth;
ae = 4 mm, ap = 1.5 mm).

4042

J. Sun, Y.B. Guo / Journal of Materials Processing Technology 209 (2009) 40364042

and stable bulk hardness (800 HK). It means surface materials


experienced signicant strain hardening in the range of 6880%
induced by surface deformation. Furthermore, the higher the cutting speed, the lower the surface hardness. This is because a
higher cutting speed generates higher temperatures which tend
to induce thermal softening for countering the dominant strain
hardening.
It should be pointed out that the rst hardness data measured
in the subsurface is slightly less than the bulk material. Some
researchers attribute this phenomenon to surface softening. The
measured surface hardness proves that the so called surface softening in literature lacks physical basis. In fact, Warren and Guo
(2006) found that the measured low microhardness value in the
subsurface is caused by edge effect on the surface during a microhardness test instead of the misinterpreted surface softening.
The explanation has been conrmed by both microhardness and
nanohardness measurement on the very top surface and in subsurface (Warren and Guo, 2006).
4. Conclusions
A series of end milling titanium Ti6Al4V experiments were
conducted to comprehensively characterize surface integrity in various milling conditions. The major conclusions may be summarized
as follows:
The milled surface shows the anisotropic nature. The simulated
surface pattern agrees with the measured one. Surface roughness value in cutting and feed directions increases with feed and
radial depth-of-cut, while it also increases in cutting direction
in the low speed range but decreases in the high speed range.
Surface roughness value in feed direction decreases with cutting
speed. The variation ranges of surface roughness values in both
directions are within 0.61.0 m.
Compressive residual normal stresses in cutting and feed directions increase with cutting speed and have a maximum around
the speed of 80 m/min. The compressive residual normal stress
in feed direction is about 30% larger than that in cutting direction. The magnitudes of residual shear stress are much smaller
than those of the residual normal stresses. Compressive residual normal stress decreases in general with feed. The highly
nonlinear coupling of mechanical and thermal loading determines the characteristics and magnitudes of residual stress
proles.
The phase seems to experiences more deformation and volume shrinkage in the near surface with the cutting speed.
However, phase transformation was not observed. The variation
trend is qualitatively consistent with that of residual stress. The
microhardness at surface is about 7090% larger than the bulk
material.

Acknowledgements
The authors would like to thank the assistance of Drs. Andrew
Warren, Mark Barkey and Mark Weaver for preparing samples and
surface integrity characterization.
References
Amin, N.A.K.M., Ismail, A.F., Khairusshima, N.M.K., 2007. Effectiveness of uncoated
WCCo and PCD inserts in end milling of titanium alloyTi6Al4V. J. Mater.
Proc. Technol. 192/193, 147158.
Canteroa, J.L., Tardiob, M.M., Cantelia, J.A., Marcosc, M., Miguelez, M.H., 2005. Dry
drilling of alloy Ti6Al4V. Int. J. Machine Tools Manuf. 45, 12461255.
Che-Haron, C.H., Jawaid, A., 2005. The effect of machining on surface integrity of
titanium alloy Ti6%Al4%V. J. Mater. Proc. Technol. 166, 188192.
Copper, D., 2006. Machining capacity, expertise must grow. SME/Aerospace Defense
Manuf. Manuf. Eng., 6365.
Ezugwu, E.O., 2005. Key improvements in the machining of difcult-to-cut aerospace
superalloys. Int. J. Machine Tools Manuf. 45, 13531367.
Ezugwu, E.O., Bonneya, J., Silvab, R.B.D., Cakir, O., 2007. Surface integrity of nished
turned Ti6Al4V alloy with PCD tools using conventional and high pressure
coolant supplies. Int. J. Machine Tools Manuf. 47, 884891.
Fitzpatrick, M.E., Fry, A.T., Holdway, P., Kandil, F.A., Shackleton, J., Suominen, L., 2002.
Determination of residual stresses by X-ray diffraction. Meas. Good Practice
Guide 52, 159.
Hauk, V., 1997. Structural and Residual Stress Analysis by Nondestructive Methods.
Elsevier Science, Danvers, MA.
Hong, H., Riga, A.T., Cahoon, J.M., 1993. Machinability of steels and titanium alloys
under lubrication. Wear 162, 3439.
Hughes, J.I., Sharman, A.R.C., Ridgway, K., 2004. The effect of tool edge preparation
on tool life and workpiece surface integrity. Proc. Inst. Mech. Eng., Part B: J. Eng.
Manuf. 218, 11131123.
Kahles, J.F., Field, M., Eylon, D., Froes, F.H., Metals, J., 1985. Mach. Titanium 37, 2735.
Li, R., Riester, L., Watkins, T.R., Blau, P.J., Shih, A.J., 2007. Metallurgical analysis and
nanoindentation characterization of Ti6Al4V workpiece and chips in highthroughput drilling. Mater. Sci. Eng. A 472, 115124.
Lopez de lacalle, L.N., Perez, J.I., Llorente, J., Sanchez, J.A., 2000. Advanced cutting
conditions for the milling of aeronautical alloys. J. Mater. Proc. Technol. 100,
111.
Norihiko, N., Akio, M., Suquru, M., 1983. Study on machining of titanium alloys. CIRP
Ann. 32, 6569.
Reissig, L., Volkl, R., Mills, M.J., Glatzel, U., 2004. Investigation of near surface structure in order to determine process-temperatures during different machining
processes of Ti6Al4V. Scr. Mater. 50, 121126.
Ribeiro, M.V., Moreira, M.R.V., Ferreira, J.R., 2003. Optimization of titanium alloy
(Ti6Al4V) machining. J. Mater. Proc. Technol. 143/144, 458463.
Snoha J., 1996. X-ray diffraction characterization of process-induced residual stress.
Army Research Laboratory ARL-TR-1204.
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., 2008. A new multi-view approach to characterize 3D chip morphology and properties in end milling titanium alloy Ti6Al4V. Int. J. Machine
Tools Manuf. 48, 14861494.
Warren, A.W., Guo, Y.B., 2006. On the clarication of surface hardening by hard
turning and grinding. Trans. NAMRI/SME 34, 309316.
Yang, X.P., Liu, C.R., Grandt, A.F., 2002. An experimental study on fatigue life variance, residual stress variance, and their correlation of face-turned and ground
Ti6Al4V samples. J. Manuf. Sci. Eng. 124, 809819.
Zlatin, N., Field, M., 1973. Procedures and precautions in machining titanium alloys.
Titanium Sci. Technol. 1, 489504.
Zoya, Z.A., Krishnamurthy, R., 2000. The performance of CBN tools in the machining
of titanium alloys. J. Mater. Proc. Technol. 100, 8086.

Você também pode gostar