You are on page 1of 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/257337215

Influence of milling on surface integrity of


Ti6Al4V-study of the metallurgical
characteristics...
Article in International Journal of Advanced Manufacturing Technology July 2012
DOI: 10.1007/s00170-012-4582-5

CITATIONS

READS

14

87

5 authors, including:
Kamel Moussaoui

Johanna Senatore

Clment Ader Institute

Paul Sabatier University - Toulouse III

8 PUBLICATIONS 24 CITATIONS

16 PUBLICATIONS 142 CITATIONS

SEE PROFILE

SEE PROFILE

Rmy Chieragatti

Frederic Monies

Institut Suprieur de l'Aronautique et de l'E

Paul Sabatier University - Toulouse III

28 PUBLICATIONS 218 CITATIONS

25 PUBLICATIONS 198 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Kamel Moussaoui


Retrieved on: 10 November 2016

Int J Adv Manuf Technol (2013) 67:14771489


DOI 10.1007/s00170-012-4582-5

ORIGINAL ARTICLE

Influence of milling on surface integrity of Ti6Al4Vstudy


of the metallurgical characteristics: microstructure
and microhardness
Kamel Moussaoui Michel Mousseigne
Johanna Senatore Rmy Chieragatti
Frdric Monies

Received: 9 July 2012 / Accepted: 23 October 2012 / Published online: 16 November 2012
Springer-Verlag London 2012

Abstract The quality of titanium alloy parts in the


aeronautical field demands high reliability, which is
largely related to surface integrity. Surface integrity
is generally defined by three parameters: a geometric
parameter, a mechanical parameter and a metallurgical
parameter. The present article addresses the influence
of milling on the metallurgical parameter for a surface
milled in Ti6Al4V material, focusing in particular on
the microstructure and microhardness. Observation of
the machined surface from a macroscopic perspective
highlight an orange peel phenomenon. This effect is
the combined result of redeposition and crushing of
the milled material. No plastically deformed layer or
lengthening of the grains was observed under the milled
surface. As far as microhardness is concerned, a slightly
softened region was observed under the milled surface. A diffusion of vanadium from the phase to
the phase was also noted, but with no change in
microstructure.
Keywords Surface integrity Ti6Al4V Milling
Microstructure Microhardness

K. Moussaoui (B) M. Mousseigne J. Senatore


R. Chieragatti F. Monies
Universit de Toulouse 3, INSA, UPS, Mines Albi, ISAE,
Toulouse, France
e-mail: moussaoui@lgmt.ups-tlse.fr
M. Mousseigne
e-mail: michel.mousseigne@univ-tlse3.fr
K. Moussaoui M. Mousseigne J. Senatore
R. Chieragatti F. Monies
ICA (Institut Clment Ader), 135 avenue de Rangueil,
31077 Toulouse, France

1 Introduction
Titanium alloys are widely used in the aeronautical
industry for applications requiring excellent mechanical
resistance at high temperatures. Furthermore, titanium
alloys have a density 40 % less than steels and outstanding resistance to corrosion. They are also utilised in the
biomedical field and the chemical industry. However,
these alloys are also known to be difficult materials
to machine. This is due to their strong resistance at
high temperature, a relatively low modulus of elasticity,
low thermal conductivity and strong chemical reactivity
[14].
The quality of titanium alloy parts in the aeronautical field requires great reliability, this being largely
dependent on surface integrity.
Surface integrity is generally defined by three parameters :
a geometric parameter : roughness
a mechanical parameter : residual stresses
a metallurgical parameter : the microstructure and
microhardness
This article adopts the perspective of a global study
into the influence of machining on the fatigue life of
aeronautical parts made of titanium alloy. The study
is divided into two parts, the first being devoted to
the influence of machining on surface integrity and the
second to the influence of surface integrity on fatigue
life. The final goal is to highlight one or more machining
criteria influencing fatigue life. The article focuses on
the influence of milling on the metallurgical parameter
of a surface machined on Ti6Al4V. The test conditions
applied are derived from an experimental design developed within the scope of the global study. By way of an

1478

introduction, a state of the art is given in what follows


immediately.
Firstly, a review of previous publications studies covering microstructural damage to the material generated
as by machining will be given.
Studies on Ti6Al4V and TI-6246 have shown that,
on a surface obtained in dry turning, a very fine plastically deformed layer formed in the sub-surface [5, 6].
This plastically deformed layer becomes thicker as
cutting speeds and feed rates increase. Using a worn
milling cutter accentuates microstructural damage and
also contributes to augmenting the thickness of the
deformed layer [6]. Che-Haron and Jawaid [5] observed
a white layer, less than 10-m thick, located just under
the machined surface. They estimate that this white
layer is due to cutter wear. Ibrahim et al. [7] also
observe this extra low interstitial phenomenon when
dry turning Ti6Al4V.
Conversely, Velasquez [8, 9], when high-speed dry
turning of Ti6Al4V, did not observe the formation of
a white layer. He showed that the preparation of the
test coupons for microscopic observations can lead to
the appearance of artefacts likely to create confusion
with the white layers. Neither did he observe phase
transformation in the machined surface even for a wide
range of cutting speeds (20660 m/min). However, he
did observe a switch in the grain direction in line
with the tool pass. He observed a layer disturbed by
turning comprising two distinct zones on the machined
surfaces. The first he referred to as zone A close to
the surface (the most deformed): here, the grains are
quasi-parallel to the cutting direction over a thickness
of between 0.5 and 5 m, and this regardless of the
cutting speed used. Zone B, just a few microns thick,
separated zone A from the core material. In zone B, the
grains were also subjected to deformation, though this
was comparatively smaller, in the direction of the cutter
pass. Increased thickness of the disturbed layer was observed at greater cutting speeds. Other researchers [10]
also observed this phenomenon during the dry milling
of Ti6Al4V, i.e., creation of a thin layer plastically
deformed.
Ulutan and Ozel [11] insist that it is essential to
study the effects of these plastic deformations as they
are significant when it comes to qualification of surface
integrity. Such deformations are caused and/or accentuated by many parameters such as cutting parameters
(cutting speed, feed rate and depth of cut), the milling
cutter parameters (rake angle, edge sharpness, shape,
coating and wear) and the parameters of the milled
workpiece (material and grain size).
Hughes et al. [12], when turning Ti6Al4V with lubrication, show that the thickness of the plastically

Int J Adv Manuf Technol (2013) 67:14771489

deformed layer increases with the depth of cut. Their


cutting conditions generated a thickness of about
12 m. They observe that the cutting speed and feed
rate, cutter geometry and state of wear do not have
any measurable effect on the thickness of the plastically deformed layer. They generally noted that the
grains close to the surface had undergone deformation
and lengthening following the cutting direction. Reissig
et al. [13] also observed this deformation in the subsurface when dry drilling Ti6Al4V. Ezugwu et al. [14]
stress that when using so-called mild cutting conditions, the shear stresses cannot cause a severe plastic
deformation in the sub-surface independently of the
lubrication conditions.
Beside deformation of the grains, microstructural
changes such as the absence of the phase close to
the surface was observed when dry drilling Ti6Al4V
[15]. Similar results were also observed by Li et al. [16]
when dry drilling Ti6Al4V. Meanwhile, Sun and Guo
[17] noted that when performing a milling operation in
good lubrication conditions, the phase close to the
surface is sensitive to deformation and that its volume
tends to diminish as cutting speed increases. They argue
that this is due to the thermal and mechanical loads
applied to the workpiece during milling. However, they
did not observe a phase transformation for the cutting
conditions used.
Studies covering the influence of machining on workpiece hardness is now considered. Che-Haron and
Jawaid [5] observe that the deformed layer leads to
microhardness greater than that of the substrate, for a
depth of up to 10 m (zone corresponding to the white
layer). Beyond this depth, microhardness diminishes
to reach a minimum at 20 m. This softening of the
sub-layer is due to very high cutting temperatures. A
hardening phenomenon takes place at 70 m. They
note that cutter flank wear affects the microstructure:
the greatest microhardness measured corresponds to
milling with a damaged cutter (this can be explained
by the fact that as cutter flank wear increases, the
greater becomes the cutter/workpiece contact surface).
They also discern that microhardness increases with
the cutting speed and, to a lesser extent, with the feed
rate. The highest microhardness value was observed at
a depth of 5 m, in a zone where the microstructure was
considerably modfied by milling.
Hughes et al. [12] observe during microhardness
measurements under the milled surface that in all
cases there is a reduction in microhardness that can
be by as much as 25 % in relation to the substrate.
This increases with depth until it reaches the level
of the substrate at a depth of about 60 m. The authors consider this softening phenomenon to be due to

Int J Adv Manuf Technol (2013) 67:14771489

1479

overaging of the titanium alloy due to the very high


cutting temperatures.
Sun and Guo [17] observe that the microhardness
on the surface is 7090 % higher than that of the core.
They show that the higher the cutting speed, the more
microhardness diminishes on the surface. Cantero et al.
[15] measure a microhardness at 50 m from the surface
30 % higher than that of the core.
Daymi et al. [18] study the influence of feed rate
on surface integrity of titanium alloy in high-speed
milling. They observed no plastic deformation in the
top layer of the machined surface. The cutting conditions influence the microhardness of the machined
surface. The major part of the changes in the microhardness was concentrated up to 140 m beneath the
surface. The microhardness increases with the increase
of feed.
The influence of machining on microstructure and
hardness will now be examined.
Fig. 1 Microstructure of the Ti6Al4V alloy studied

2 Experimental details
The material studied was an + two-phase titanium
alloy, Ti6Al4V. Its chemical composition in accordance
with standard AFNOR L14-601 and its microstructure
are shown, respectively, in Table 1 and Fig. 1, representing a bimodal or duplex + microstructure. The
phase is shown in black and the phase white. Table 2
shows its mechanical properties and hardness that was
measured in the laboratory.
The core hardness was averaged out over 120 measurements that were made at the core of each sample
(Table 3).
Machining was conducted on a HURCO brand
three-axis milling machine with 10-kW power. The cutter used was of the ISCAR brand with reference ADKT
150524R-HM IC928. This is a PVD TiAlN-coated carbide cutter. The geometry and cutter dimensions are
shown in Fig. 2.
The machining conditions are given in Table 4. The
cutting parameters are given by Airbus Toulouse. Each
test sample was machined with a new cutter. The aim
is to assess the influence of cutting conditions on the
metallurgical appearance independently of cutter wear.
The cutting conditions used in this article have been
identified from test E1. It was arbitrarily chosen from
the experimental design of the global study. It includes

a finishing pass in order to observe whether or not its


influence on surface integrity. Only the cutting speed
and depth of cut in roughing vary. All tests were performed dry, with a feed rate of 0.133 mm/rev, taken
in the global study experimental design, and a corner
radius R of 2.5 mm.
In order to characterise the microstructure of the
milled surface, metallographic observation was resorted to. This involved observing mainly the plastic deformation of the superficial layer microstructure affected by machining. All the present observations were made using a scanning electron microscope
(SEM).
The sampling procedure observed involved using a
Buhler grinder with a velocity of rotation 2,500 rpm, a
feed rate 1.2 mm/min and with good lubrication conditions so as to reduce thermal effects. The coupons
were polished mechanically with sandpaper with grain
size decreasing from 800 to 4,000 grains/mm2 . They
were then chemically etched in a bath comprising 3 %
HNO3 , 4 % HF and 93 % H2 O. This type of etching
is used for observations in optical microscopy or scan
electronic microscopy in secondary electron (SE) observation mode. Etching time was about 15 s.

Table 1 Chemical composition of the titanium alloy studied (AFNOR L14-601)


Element

Ti

Al

Fe

% weight

Base

5.56.75

3.54.5

<0.25

<0.2

<0.08

<0.05

<0.01

1480

Int J Adv Manuf Technol (2013) 67:14771489

Table 2 Mechanical properties


Rm (MPa)

Rp0.2
(MPa)

A (%)

K IC

(MPa m)

[19]
(g cm3 )

[19]
(W m1 K1 )

Hardness
measured (HV)

9001,160

830

810

39

4.43

335.65

Table 3 Core hardness of the studied material: Ti6Al4V


Mean

SD

Max

Min

335.65 HV

19.24

402.5 HV

305.8 HV

Microhardness measurements are frequently used to


characterise the metallurgical condition of machined
surfaces. This approach allows the densities of dislocations in the microstructure to be quantified. It was
decided to characterise microhardness using Vickers
testing. The advantage of using a pyramid penetrator
(as with a conical penetrator) is that the similarity
law is automatically respected: when you change the
load, you obtain geometrically similar imprints and,
thus, identical values for hardness. Furthermore, the
advantage with diamond is the absence of deformation
of the penetrator when measuring high hardness values.
Vickers testing is equally suited to both very hard materials and soft materials as, due to the constancy of the
angle of penetration, the measurement remains independent of the load. As the marks are relatively small,
as near perfect polishing as possible is desirable [20].
The measurements were made using a MITUTUYO
brand MVK-H1 type microhardness instrument with a
Vickers indenter and force F=300 g for 20 s.
Microhardness measurements were made in zone 1,
perpendicular to the long transverse direction TL of
the material, in order to measure change in hardness
in relation to depth and then in zone 2 to determine the
core hardness for each sample (Fig. 3a). In zone 1, each
value was averaged out over three points distant from
each other by 200 m along the rolling direction L and
30 m along the TC (short transverse) direction. The
measurement points were distributed in an alternating
matrix 3 12 and the first line of measurements was

Fig. 2 ISCAR insert dimensions and geometry

made 35 m from the machined surface. The details


for zone 1 are shown in Fig. 3b. In zone 2, the points
were distant from each other 200 m along L and
200 m along TC. They were distributed in a conventional matrix 3 10 that is 30 measurements to the
core for each sample. Measurements on the machined
surface were also made. They were averaged over nine
measurements and thus represent the hardness values
at a depth of 0 m on the graphs.
The chemical pickling is carried out in a bath of 2 %
HF, 20 % HNO3 and 78 % H2 O. This was done at ambient temperature and without stirring. This pickling was
performed to see if there was a work-hardened layer
effect on mechanical polishing on the microhardness.

3 Results and discussion


3.1 Analysis of the milled surfaces
Two principal phenomena were observed on all the
machined surfaces: traces of feed and deformations
perpendicular to the feed direction generated by the
cutter (Fig. 4) and an orange peel texture observable to
the naked eye. This textures presence can be explained
mainly by a redeposition and crushing of the material
during milling, micro-cavities and chips that adhered to
the machined surface.
The deformations due to feed (Fig. 4b and c) are
the result of plastic flow of the material during cutting
[14, 21] but can also result from too great an edge sharpness and corner radius. The micro-cavities, meanwhile,
are generated by brittle fracture of carbide inclusions
on the surface (Fig. 4f) during shearing of the material
by the cutter [14].

Int J Adv Manuf Technol (2013) 67:14771489

1481

Table 4 Test parameters


Sample

Cutting speed
(m/min)

Depth of cut
roughing
(apeb ) (mm)

Depth of cut
finishing
(ap f in ) (mm)

E1
E2
E3
E4
E5

54
54
54
78
78

1.65
1.65
2.35
1.65
2.35

0.15

The redeposition of material was observed on all


the surfaces machine-dried but can also emerge with
conventional lubrication [14]. According to Ezugwu
[14], this phenomenon is due to fine chip particles
that get welded to the machined surface to form a
composite surface. Indeed, machining titanium can generate high local temperatures that will facilitate the
welding/adherence process. The temperature at the cutter/chip interface when turning a titanium alloy with a
cutting speed of 200 m/min and feed rate of 0.1 mm/rev
using a carbide cutter can reach 1,100 dry and 900 C
with conventional lubrication [22]. This author also
states that cutter wear can also intervene in this phenomenon. In the present study, as each surface was
milled with a new mill, this proves that cutter wear is
not what triggers the phenomenon though it does tend
to accentuate it.
In the present context, this redeposition of material is revealed to the naked eye by an orange peel
phenomenon on the machined surfaces. As observed
by Ezugwu [14], fine chip particles (Fig. 4c, d) could
be seen adhering to the machined surface. But the

orange peel texture comes mainly from feed marks


(Fig. 4f). The latter are caused during machining by
the stretching of strips of material until they suddenly
rupture (Fig. 4f).
Following this first macroscopic observation of the
surface, microscopic observation under the machined
surface was conducted.
Figure 5ae represent sections perpendicular to the
machined surfaces and thus to the feed direction. No
obvious defects can be observed under the surface: no
plastic deformations and no lengthening of the grains
in the feed direction. However, the latter have already
been observed in titanium alloys, as with Velasquez
[8] who observed such phenomena when dry turning
Ti6Al4V for different cutting conditions. He noted that
for a cutting speed of between 20 and 100 m/min, the
thickness of the plastically deformed layer (lengthening
of grains) varied between 0.5 and 1.5 m and reached
4.25 m for 420 m/min. In his view, this zone has microstructural characteristics that recall the shear bands
seen in the chips.
It can thus be concluded that severe cutting conditions are required to generate this layer. As far as the
present study is concerned, no clear evidence was observed of a plastically deformed layer and lengthening
of the grains due to the fact that the cutting conditions
were less severe.

3.2 Influence of machining on hardness


Considering the literature published on the subject, it emerges that a fine work hardened layer

Fig. 3 Microhardness
measurement protocol

(a) Definition of measurement protocol

(b) Detail of zone 1

1482

Int J Adv Manuf Technol (2013) 67:14771489

Fig. 4 Surfaces milled dry

(a) E2 (V=54 m/min, apeb=1.65 mm)

(b) E3 (V=54 m/min, apeb=2.35 mm)

(c) E4 (V=78 m/min, apeb=1.65 mm)

(d) E5 (V=78 m/min, apeb=2.35 mm)

(e) E1 (V=54 m/min, apfin=0.15 mm)

(f) E1 surface details

(from a few micrometers to several tens of micrometers) may appear during machining. As a first
approach, the nanohardness measurement technique
appears to be the most appropriate. However, considering the grain size of around 20 m and the twophase character of the material studied, the results
obtained may show considerable dispersion. In this
case, the microhardness technique may give better
results.
The following paragraphs will provide a study of
these two measurement methods and discuss their relative merits.

3.2.1 Nanohardness measurement


In order to characterise the presence of a work
hardened layer on the surface, nanoindentation or
nanohardness tests were performed. This method allows the hardness of very fine layers or that of grains
one by one to be measured for a two-phase material.
To this purpose, a CSM Instruments machine was used
with a Berkovich indenter. The depth of penetration
was 1,000 nm with a 800 mN load for 10 s.
The surfaces were prepared in accordance with the
method described in Section 2, but an additional stage

Int J Adv Manuf Technol (2013) 67:14771489

1483

Fig. 5 Observation of the


microstructure under the
machined surface

(a) E2 (V=54 m/min, apeb=1.65 mm)

(b) E3 (V=54 m/min, apeb=2.35 mm)

(c) E4 (V=78 m/min, apeb=1.65 mm)

(d) E5 (V=78 m/min, apeb=2.35 mm)

(e) E1 (V=54 m/min, apfin=0.15 mm)

was introduced for the nanohardness measurement involving mirror polishing using a 3-m diamond paste.
Ten measurements were made on each milled surface
to characterise nanohardness.
Figure 6 shows a very wide dispersion of nanohardness results. This dispersion can be explained by the
fact that a two-phase material is involved in the study,
and indenting can thus be applied to or , a grain
boundary or a defect.
This induced a change in the hardness measurement
method in favour of microhardness testing so as to reduce that phenomenons influence. The measurement
protocol and the effects of the indentation load and

polishing on the hardness results will be described in


what follows.
3.2.2 Analysis ways of measurement
This paragraph addresses the issue of setting up a microhardness measurement protocol. This preliminary
study proved necessary due to the first microhardness
results being extremely dispersed; it appeared important to see the influence of measurement and the surface preparation on the results. First of all, influence of
the force of indentation on microhardness was studied
and then the influence of polishing on microhardness.

1484

Int J Adv Manuf Technol (2013) 67:14771489

Fig. 6 Hardness and standard deviation in surface measurements of nanohardness

All measurements were made at the core of the material to limit the influence of additional parameters like
machining.
Inf luence of the indentation force The loads tested
were as follows: 1,000, 500, 300, 200, 100 and 50. The
unit was gramme-force. For a given load, 30 mea-

Fig. 7 Influence of the


indentation force on
microhardness

surements were performed. The tests were done on a


mechanically polished sample. These tests sought to
highlight the difficulty of measuring Vickers microhardness precisely in the case of TA6V and determine an
adequate force combining a relatively low dispersion
with a relatively small indentation mark so as to be able
to make measurements close to the machined surface.

Int J Adv Manuf Technol (2013) 67:14771489

1485

Table 5 Length of the diagonals and depth of penetration


Indentation
force (F) (gf)

D measured
m

Depth calculated
(m)

1,000
500
300
200
100
50

74.3
53
40.5
33
23.2
16.2

10.6
7.6
5.8
4.7
3.3
2.3

Figure 7 shows the standard deviation for the


different loads tested. Table 5 shows the length of the
diagonal D of the imprint (mean of the two diagonals
D1 and D2 Fig. 8) and the depth of penetration of the
indenter for each load tested. The indenter is a diamond
pyramid with a square base and the angle of the two
opposite sides is 136 (Fig. 8) [20].
In Fig. 7, it can be seen that the smaller the load is,
the more the standard deviation increases. This derives
from the microstructure of the material involved. Indeed, the said material was an + two-phase titanium with a phase harder than the phase [23]. In
Table 5, it can be seen that the smaller the load, the
smaller will be the length D and depth of penetration.
As the grain size is about 20 m, with a low load, there
is a tendency to indent a single grain rather than several
of them. This leads to a greater dispersion according
to whether indenting occurs in phase or and the
proportion of the latter.
The present study sought to observe the changes
in hardness going from the machined surface down
to the substrate. This impacts the choice of force of
indentation as the aim is to indent extremely close to
the surface. It was therefore decided to choose a force
of indentation F=300 gf as, firstly, it has a relatively
low standard deviation compared with the other forces,

and, secondly, it allows measurements to be started


fairly close to the surface, at about 35 m.
Inf luence of polishing In most articles addressing microhardness, a test sample polishing phase is presented
without any analysis of its influence being proposed. It
is, thus, proposed to study the influence of polishing
on the microhardness results; to do so, three types of
samples were prepared:
a mechanically polished sample, referred to as PM;
a sample mechanically polished then pickled with
a loss in thickness of about 2.71 m, referred to as
PM+De1;
a sample mechanically polished then pickled with
a loss in thickness of about 4.73 m, referred to as
PM+De2.
The tests were performed using a force F=300 gf and
each result was averaged out over 20 measurements.
The results are shown in Table 6 and Fig. 9.
It can be seen that the hardness is pretty similar for
the different types of surface preparation. This suggests
that the work-hardened layer does not interfere with
the microhardness measurements at least for the load
used in the present study. All the measurements presented in what follows in this article were made on
mechanically polished surfaces.
3.3 Analysis of measurement set-up for the study
of an + two-phase titanium
Figure 10ac show microhardness curves in relation to
distance under the machined surface. The points at zero
were measured on the machined surface.
The first observation made on Fig. 10 relates
to significant dispersion of the measurements taken.
The curves oscillate enormously even though each
point plotted was averaged over several measurements
(Fig. 11). The scale of dispersion over the different tests
performed (highlighted in Fig. 11) can be explained
by the fact that the material studied is an + twophase titanium alloy with a phase harder than the
phase. Despite the prior study conducted into the
choice of indenter and the force of indentation to
minimise measurement dispersion, it can be seen that

Table 6 Average hardness

Fig. 8 Imprint of the indenter

Sample

Mean (HV)

PM
PM+De1
PM+De2

336.46
333.75
339.14

1486

Int J Adv Manuf Technol (2013) 67:14771489

Fig. 9 Influence of polishing


on microhardness

the size of the indenter will always be insufficient in


relation to the grain size of the alloy. There will always
be a tendency to indent over few grains, which will in
turn inevitably lead to dispersion. This first observation
leads to concluding that the test facilities used for the
hardness measurements are not suited to the chosen
alloy.
Nevertheless, despite the level of dispersions, a softening phenomenon can be observed in the sub-surface
at varying depths (Figs. 10 and 12) that can be seen on
each of the charts plotted. A number of researchers
[5, 12] have evoked this material softening phenomenon in the sub-surface and give high cutting temperatures as an explanation.
The alloys composition at the core and close to the
surface was determined using an EDX analyser coupled
with an SEM, and this led to the following observation:
the proportion of vanadium is 11.6 % in the phase
and 2.4 % in the phase in the core,
the proportion of vanadium is 3.7 % in the phase
and 3.9 % in the phase under the machined surface (max. depth about 78 m).
What can therefore be seen is a diffusion of vanadium from the phase to the phase. The diffusion of
vanadium into the phase is less signficant as it is more
present in the Ti6Al4V alloy than the phase. Other
researchers have also observed the diffusion of [13, 16].
It should be noted that this diffusion occurred without
any change to the microstructure observable to our
instruments. This diffusion seems to be the origin of the
softening phenomenon observed under the machined
surface. Researchers have shown that the greater the

proportion of vanadium in the phase, the greater will


be the hardness [2426].

4 Conclusion
The central topic addressed by the present article can
be summarised as an investigation into the influence
of machining parameters on Ti6Al4V machined surface integrity with a focus on how machining processes
influence the metallurgy of the titanium alloy studied.
Observing the surface from a macroscopic perspective
revealed an orange peel phenomenon over all the machined surfaces. This effect results from the combined
redeposition (fine chip particles adhering to the surface
as a result of the high temperatures generated during
cutting) and a crushing of the machined material then a
stretching and sudden rupture of some strips of material
caused by the high temperatures generated when dry
machining.
Microscopic observation of the surface was then
performed to seek evidence of a plastically deformed
layer and grains deformed in the feed direction. This
phenomenon is mentioned in many articles covering
the machining of titanium alloys but failed to emerge on
the materials chosen in the present series of tests. This
tends to prove that the phenomenon is only triggered
during tests that apply severe cutting conditions.
To conclude, the variations in hardness going down
from the machined surface into the core of the workpiece were considered. The main conclusion is that
conventional hardness measurement instruments are
not suited to the two-phase titanium alloy studied as
the very nature of this alloy is inducive to considerable

Int J Adv Manuf Technol (2013) 67:14771489

1487

Fig. 10 Evolution of
microhardness under the
machined surface

(a) V=54 m/min, f=0.133 mm/rev

(b) V=78 m/min, f=0.133 mm/rev

(c) V=54 m/min, f=0.133 mm/rev, E1 : roughing (apeb=1.65 mm)


+ finishing (apfin=0.15 mm), E2 : roughing (apeb=1.65 mm)

1488

Int J Adv Manuf Technol (2013) 67:14771489

Fig. 11 Dispersion observed


on microhardness
measurements,
V = 54 m/min,
f = 0.133 mm/rev, E1 :
roughing (apeb = 1.65 mm) +
finishing (ap f in = 0.15 mm),
E2 : roughing
(apeb = 1.65 mm)

dispersion in measurement values. Nevertheless, the


presence of a softened zone under the surface was highlighted. This softening is due to the high temperatures
generated during cutting resulting from Ti6Al4Vs low
thermal conductivity. These temperatures also provided the conditions for a diffusion of vanadium from
the phase of the alloy to the phase under the
machined surface though without changes occurring in
the microstructure. It is likely that this diffusion led to
the softening phenomenon encountered under the machined surface. The transmission electron microscope
and/or definition of work hardening by the width of
the diffraction peak were the methods available to

Fig. 12 Softening observed in


the sub-surface

highlight this work hardened layer. Indeed, the width of


the diffraction peak can be related to the rate of dislocations in the volume of material analysed. This rate of
dislocation is also directly related to the microhardness
of the material.
In view of the results on the microstructure and
microhardness, we can say that the roughing affects
the machined surface very little. It is the last step,
namely the finishing pass, which defines the thermomechanical state of the surface. This leads us to say
that the roughing can be machined with more severe
cutting conditions and could lead to increased productivity. However, to confirm our hypothesis, a study on

Int J Adv Manuf Technol (2013) 67:14771489

residual stresses and roughness affected by the cutting


is in progress and will be presented in a forthcoming
publication.
Acknowledgement The authors would like to acknowledge
Airbus Toulouse for providing the material and tools in this work.

References
1. Ezugwu EO 2005 Key improvements in the machining
of difficult-to-cut aerospace superalloys. Int J Mach Tools
Manuf 45(1213):13531367
2. Konig W (1978) Applied research on the machinability of
titanium and its alloys. In: Proceedings of 47th meeting
of AGARD structural and materials panel. Florence, Italy,
pp 1.11.10
3. Sun J, Guo YB (2008) A new multi-view approach to characterize 3D chip morphology and properties in end milling
titanium Ti-6Al-4V. Int J Mach Tools Manuf 48(1213):1486
1494
4. Ezugwu EO, Wang ZM (1997) Titanium alloys and their
machinabilitya review. J Mater Process Technol 68(3):262
274
5. Che-Haron CH, Jawaid A (2005) The effect of machining on
surface integrity of titanium alloy TI-6 % AL-4 % V. J Mater
Process Technol 166(2):188192
6. Che-Haron CH (2001) Tool life and surface integrity in turning titanium alloy. J Mater Process Technol 118(13):231237
7. Ibrahim GA, Che-Haron CH, Ghani JA (2009) The effect of
dry machining on surface integrity of titanium alloy Ti-6Al4V ELI. J Appl Sci 9:121127
8. Puerta Velasquez JD (2007) Etude des copeaux et de
lintgrit de surface en usinage grande vitesse de lalliage
de titane TA6V. PhD thesis, Universit Paul Verlaine, Metz
9. Puerta Velasquez JD, Tidu A, Bolle B, Chevrier P, Fundenberger J-J (2010) Sub-surface and surface analysis of
high speed machined Ti-6Al-4V alloy. Mater Sci Eng, A
527(1011):25722578
10. Mhamdi M-B, Boujelbene M, Bayraktar E, Zghal A (2012)
Surface integrity of titanium alloy Ti-6Al-4V in ball end
milling. Phys Procedia 25(0):355362
11. Ulutan D, Ozel T (2011) Machining induced surface integrity
in titanium and nickel alloys: a review. Int J Mach Tools
Manuf 51(3):250280

1489
12. Hughes JI, Sharman ARC, Ridgway K (2006) The effect of
cutting tool material and edge geometry on tool life and
workpiece surface integrity. Proc Inst Mech Eng, B J Eng
Manuf 220(2):93107
13. Reissig L, Vlkl R, Mills MJ, Glatzel U (2004) Investigation of near surface structure in order to determine
process-temperatures during different machining processes
of Ti6Al4V. Scr Mater 50(1):121126
14. Ezugwu EO, Bonney J, Da Silva RB, Cakir O (2007) Surface
integrity of finished turned Ti-6Al-4V alloy with PCD tools
using conventional and high pressure coolant supplies. Int J
Mach Tools Manuf 47(6):884891
15. Cantero JL, Tardio MM, Canteli JA, Marcos M, Migulez
MH (2005) Dry drilling of alloy Ti-6Al-4V. Int J Mach Tools
Manuf 45(11):12461255
16. Li R, Riester L, Watkins TR, Blau PJ, Shih AJ (2008) Metallurgical analysis and nanoindentation characterization of
Ti-6Al-4V workpiece and chips in high-throughput drilling.
Mater Sci Eng, A 472(12):115124
17. Sun J, Guo YB (2009) A comprehensive experimental study
on surface integrity by end milling Ti-6Al-4V. J Mater
Process Technol 209(8):40364042
18. Daymi A, Boujelbene M, Bayraktar E, Ben Amara A,
Katundi D (2011) Influence of feed rate on surface integrity
of titanium alloy in high speed milling. Adv Mat Res 264
265:12281233
19. Leyens C, Peters M (2003) Titanium and titanium alloys:
fundamentals and applications. Wiley, Weinheim
20. Franois D (2005) Essais mcaniques des mtaux-essais de
duret. Techniques de lingnieur, M4160, p 10
21. Zhou L, Shimizu J, Muroya A, Eda H (2003) Material removal mechanism beyond plastic wave propagation rate.
Precis Eng 27(2):109116
22. Kitagawa T, Kubo A, Maekawa K (1997) Temperature and
wear of cutting tools in high-speed machining of inconel 718
and Ti-6Al-6V-2Sn. Wear 202(2):142148
23. Ankem S, Margolin H (1986) Modeling deformation in 2phase alloys. J Met 38:2529
24. Welsch G, Boyer R, Collings EW (1994) Materials properties
handbook: titanium alloys. ASM International
25. Banerjee R, Collins PC, Bhattacharyya D, Banerjee S, Fraser
HL (2003) Microstructural evolution in laser deposited
compositionally graded \ titanium-vanadium alloys. Acta
Mater 51(11):32773292
26. Craighead CM, Simmons OW, Eastwood LW (1950)
Titanium binary alloys. Transactions AIME, Trans AIME
188

You might also like