Você está na página 1de 10

Journal of Materials Processing Technology 215 (2015) 95104

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

Progressive tool-wear in machining of room-temperature austenitic


NiTi alloys: The inuence of cooling/lubricating, melting, and heat
treatment conditions
Y. Kaynak a, , S.W. Robertson b , H.E. Karaca c,d , I.S. Jawahir c,d
a

Department of Mechanical Engineering, Faculty of Technology, Marmara University, Goztepe Campus, Kadikoy, 34722 Istanbul, Turkey
Fathom Engineering, Berkeley, CA 94710, USA
c
Institute for Sustainable Manufacturing (ISM), University of Kentucky, Lexington, KY 40506, USA
d
Department of Mechanical Engineering, University of Kentucky, Lexington, KY 40506, USA
b

a r t i c l e

i n f o

Article history:
Received 5 March 2014
Received in revised form 14 July 2014
Accepted 16 July 2014
Available online 24 July 2014
Keywords:
Machining
Tool-wear
Room-temperature austenitic NiTi alloys
Cooling/lubricating
Melting
Heat treatment

a b s t r a c t
This study investigates the effects of fabrication (vacuum arc remelting (VAR) melted vs. vacuum
induction melting combined with VAR (VIM + VAR)), processing (hot rolled + fully annealed vs. cold
worked + superelastic anneal) and machining conditions (dry, cryogenic, and minimum quantity lubrication (MQL)) of NiTi alloys on their progressive tool-wear behavior. Experimental ndings reveal that
cryogenic machining substantially improves the performance of cutting tools by reducing the progressive
tool-wear in machining of the room-temperature austenitic NiTi alloys. Therefore, cryogenic machining could result in improved productivity and reduced manufacturing costs compared to dry and MQL
machining. Experimental evidence suggests that cold working did not alter the progressive tool-wear substantially; however, the presence of carbide inclusions increased the progressive tool-wear in machining
NiTi. Surface quality of machined samples under cryogenic machining presents promising improvement
upon short-duration machining compared to dry and MQL machining, but all three techniques resulted
in comparable quality after 4 min of machining.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Although shape memory and superelastic alloys are used in
different industries, their use in biomedical devices is their most
common application (Morgan, 2004). Nitinol is an alloy containing approximately 50 atomic % nickel and 50 atomic % titanium,
and small compositional changes around this 50:50 ratio drastically changes its operating characteristics (Morgan, 2004). Slightly
nickel-rich NiTi alloys have a unique behavior known as superleasticity near body temperature, which is utilized in the vast
majority of medical applications (Duerig et al., 1999). They can be
used as orthopedic implants, cardiovascular devices, and surgical
instruments, as well as orthodontic devices and endodontic les
(Machado and Savi, 2003). Machining of NiTi alloys is extremely difcult in the fabrication of their devices and components even when

Corresponding author. Tel.: +90 216 336 5770; fax: +90 216 337 89 87.
E-mail addresses: yusuf kaynak@yahoo.com, yusuf.kaynak@marmara.edu.tr
(Y. Kaynak).
http://dx.doi.org/10.1016/j.jmatprotec.2014.07.015
0924-0136/ 2014 Elsevier B.V. All rights reserved.

optimized conditions, adopted from other difcult-to-cut Ni and Ti


based alloys, are employed, due to the superelasticity (Kong et al.,
2011), work hardening (Kong et al., 2013), temperature and stress
induced phase transformation and ductility (Weinert and Petzoldt,
2004). The challenges of machining of this unique alloy have limited
its adoption in the orthopedics industry which relies heavily on
turning and milling operations for manufacturing of macro-scale
products.
When Nitinol is machined, manufacturers may select a very low
cutting speed, depth of cut, and feed rate which reduce the productivity and increase the cost of processing signicantly. Researchers
already reported that selected cutting parameters in machining
NiTi alloy have large effect on controlling tool-wear. Weinert
and Petzoldt (2004) reported that selecting 100 m/min cutting
speed helps to reduce tool-wear and cutting forces in machining
Ni50.9 Ti49.1 alloy. Weinert et al. (2004) optimized cutting parameters for different production steps including drilling, external
turning, internal turning, etc., to improve productivity. Additionally, they reported that uncoated carbide tools are not suitable
for machining NiTi alloys and recommended using a multi-layer

96

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

Fig. 1. Cross-section microstructure of room temperature austenitic NiTi.

coated cemented carbide to reduce tool-wear and hence to improve


machining performance in turning process of NiTi alloy (Weinert
et al., 2004). The temperature-dependent mechanical behavior of
NiTi alloys, machining conditions (cooling, lubricating, etc.) must be
considered to determine their potential contribution to improving
machinability of this alloy.
The positive effects of cryogenic cooling on the progressive tool-wear, tool-life, machining performance (Kaynak, 2014a)
and surface integrity have been reported in machining of
Inconel 718 (Pusavec et al., 2011), Ti6Al4V (Bermingham et al.,
2011), steels (Dhar et al., 2001, 2002), Ti-based aerospace alloys
(Klocke et al., 2013), and various engineering materials (Kaynak
et al., 2014a). Since the properties of NiTi alloys are highly
temperature-dependent, it is reasonable to hypothesize that cryogenic machining may inuence, and perhaps can improve the
machinability of this alloy. Indeed, a few recent studies have
demonstrated the positive contribution of cooling, particularly
cryogenic cooling, during machining on the tool-wear (Kaynak
et al., 2013a), surface quality of room-temperature martensitic NiTi
alloys (Kaynak et al., 2013b), and tool-wear rate and surface quality
of room-temperature austenitic NiTi alloy at various cutting speeds
(Kaynak, 2014b).
In this current study, we investigated the effects of various
machining conditions (dry, MQL, and cryogenic) on the progressive
tool-wear of room-temperature-austenitic NiTi alloys. Furthermore, the effects of melting techniques (VAR vs. VIM + VAR), prior
processing (cold worked vs. hot rolled) and heat treatments on
the tool performance were evaluated on the room-temperatureaustenitic NiTi alloys. To evaluate the product performance induced
from processing, the surface quality and topography of the
machined surfaces were considered.

2. Experimental procedure
2.1. Work materials
In this study, three selected room-temperature austenitic NiTi
(50.8 at % Ni) alloys were produced as round bars of 12.5 mm
diameter. These alloys were designated I, II, and III with details
provided in Table 1 and Fig. 1. Alloys I and II were fabricated by
using vacuum arc remelting processes, while alloy III was fabricated
by vacuum induction melting combined with VAR. The VAR process imparts inclusions of the Ti4 Ni2 O variety (oxide inclusions)
due to oxygen impurities from the raw titanium and/or oxygen inltration from an inadequate vacuum. VIM + VAR processing
imparts both oxide inclusions and TiC inclusions (carbide inclusions). The carbides form due to the interaction of the molten
Nitinol with the VIM graphite crucible. Differences in the morphology of inclusions (carbides are generally cuboidal, whereas
oxides are typically spheroidal or rod-like) and their elastic moduli
(carbides are 230 GPa vs. oxides 176 GPa) have been reported previously and may contribute to tool-wear characteristics (Saffari

Table 1
Melting conditions, processing conditions, heat treatment and phase transformation
temperatures (martensite start (Ms ), martensite nish (Mf ), austenite start (As ) and
austenite nish (Af )) in C of selected alloys.
Alloy

Melt

Work

Heat treatment

Ms

Mf

As

I
II
III

VAR
VAR
VIM + VAR

CW
HR
HR

Superelastic
Full anneal
Full anneal

28
32
24

44
71
36

16
43
11

Af
3
11
1

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

et al., 2013). In addition to fabricating techniques, the effects of


prior mechanical working of the NiTi bars fabricated by VAR were
studied. Specically, the two common conditions were investigated: (a) cold-worked to 30% area reduction followed by a
superelastic heat treatment (500 C for 5 min), and (b) hot-rolled
followed by a full anneal (750 C for 30 min). The cold working
(CW) plus superelastic treatment retains a signicant dislocation
density in the microstructure that results in reduced ductility
(<20% elongation to fracture) which represents the most common
treatment that is utilized for NiTi biomedical devices. Conversely,
the hot rolled plus full anneal (HR) condition eliminates all prior
cold work, decreases dislocation density and exhibits >50% ductility. In general, this condition is used as an intermediate product
form, not for nished NiTi components. The phase transformation temperatures were determined from differential scanning
calorimetry (DSC) peaks using the slope line extension method.To
examine cross-section microstructure of as received room temperature austenitic NiTi alloys (Alloys I, II, and III), specimens
were cold-mounted in cross-section, ground and polished using
conventional techniques, and etched using a solution of 3.2 vol.%
HF + 14.6 vol.% HNO3 + 82.2 vol.% H2 O. The microstructure of these
as received specimens was examined by digital optical microscopy,
as shown in Fig. 1. It has to be underlined that the martensitic laths
visible in the micrographs (Fig. 1) are an artifact from the mechanical polishing stresses. All three alloys are austenitic at room
temperature.
2.2. Machining parameters, cooling/lubricating and
measurements
A DCGT11T308HP-grade KC5410 cutting tool insert with TiB2
coating was used in the experiments. The edge radius of the tools
varied between 18 and 20 m. This selected insert with coating is
the most commonly used tool type in industry for turning of NiTi
shape memory alloys (Kaynak et al., 2013b). The tool holder was
an SDJCL123BH5M. The cutting experiments were conducted on
a Mazak QT-10 CNC Turning Center. During the machining tests,
based on our experience and recommendation from industry, a constant feed rate, f = 0.05 mm/rev, constant depth of cut, d = 0.5 mm,
and constant cutting speed, V = 25 m/min, was used. The cryogenic
coolant was liquid nitrogen, applied under 1.5 MPa pressure, and
approximately 10 g/s mass ow rate. Liquid nitrogen was delivered
to the cutting region through two nozzles each with a 4.78 mm
diameter. One of them was placed to the rake face of cutting tool;
while other was placed to the back of the tool holder to deliver
liquid nitrogen toward to cutting tool tip at an approximately 55
angle relative to the rake face of cutting tool, as shown in Fig. 2.
For minimum quantity lubrication (MQL), a UNIST system with
Coolube 2210EP metalworking lubricant was used at a ow rate of
60 ml/h, and approximately under 0.4 MPa air pressure. MQL was
applied at the rake face region of the tool. Tool-wear was periodically measured during the progressive tool-wear experiments
using an optical microscope after every minute of machining operation. Surface roughness of the workpiece was measured using a
white light interferometry optical proler ZYGO 3D New View
7300. Surface roughness, Ra , of each sample was determined by
taking the average of ve measurements.
3. Experimental results and discussion
3.1. Progressive tool-wear
Tool-wear is the undesirable result of the elevated frictional
forces. Although low thermal conductivity and work hardening
are the major reasons for high tool-wear in difcult-to-machine

97

Fig. 2. Liquid nitrogen delivery system through two nozzles placed at the rake and
ank face of the cutting tool.

materials such as Inconel 718 and Ti6Al4V alloy, phase transformation, superelasticity, microstructural purity, low elasticity
modulus and limited ductility of NiTi alloys are potential additional
contributors. Since the depth of cut is normally less than the radius
of the tool nose in the nishing operations, the maximum ank
wear land width measured at the major ank face is not a suitable
criterion for evaluating the progressive tool-wear. Instead, the ank
wear land in the nose area, according to ISO, represented by VBC is a
more meaningful parameter for assessing tool-wear (Shahabi and
Ratnam, 2009). Therefore, the maximum ank wear at nose area
(VBCmax ) was used to characterize the progressive-tool wear in this
study.
3.1.1. Alloy I: VAR-melted, cold-worked + superelastic heat
treatment
In machining of Alloy I, extremely rapid tool-wear was observed
within a short cutting time (1 min) in dry and MQL conditions, however, tool-wear rate was much lower in cryogenic machining, as
shown in Fig. 3. In dry and MQL conditions, notch wear at the depth
of cut line on the nose region was the dominating wear pattern over
time. Conversely, ank wear at the nose region was observed over
time in cryogenic machining. The benet of cryogenic machining
was the elimination of notch wear on the nose region and substantially reducing progressive tool-wear over time in comparison with
other conditions (dry and MQL). The quantitative measurements of
wear in dry, MQL, and cryogenic machining with respect to cutting
time are shown in Fig. 4.
As shown in Fig. 4, the tool-wear after 4 min of cutting was
approximately 200 m, 440 m and 630 m in the cryogenic, dry
and MQL machining, respectively. MQL was expected to reduce the
progressive-tool wear as it lubricates the cutting region, primarily the tool-chip interface. However, maximum tool-wear rate was
observed in MQL machining after 4 min cutting process. In particular, a sharp increase in tool-wear was observed between three to
4 min of cutting, and this attributed to the chipping (aking) on
the nose region. Grooves are observed to be parallel to each other
throughout the nose region and they are the dominant abrasive
wear mechanism. Similar trend was also observed for the other
two NiTi alloys.
Fig. 5 illustrates the wear phenomena on the rake face of cutting
tools after 4 min of cutting in all three conditions. Cryogenic cooling helps to improve the tool performance substantially; however,

98

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

Fig. 3. Images of the tool-wear progression at the nose region in machining of NiTi Alloy I.

different wear mechanisms such as adhesion, chipping, and failure


were observed in MQL and dry machining process.
Chipping on the rake face of the cutting tool used in dry machining was evident. Additionally, chip-to-tool welding (chip debris
attached to the cutting edges) on the main cutting edge of the cutting tool was also observed (Fig. 6). Fracture on both the major and
minor cutting edges of cutting tool used in MQL machining was
observed as shown in Fig. 5. In cryogenic machining, small grooves
on the rake face of cutting tool can be seen in Fig. 5. However, there

Fig. 4. The progression of maximum ank-wear at nose region with time under dry,
MQL, and cryogenic machining of NiTi Alloy I.

were no observed fractures, chipping, or adhesion during the in


cryogenic machining of NiTi Alloy I.

3.1.2. Alloy II: VAR-melted, hot-rolled + full anneal


In Alloys II, the greatest wear was observed in dry machining, while signicant improvement in tool-life was achieved by
cryogenic machining, as shown in Fig. 7. It is clear that Alloys I
and II showed similar wear progression. MQL shows uniform wear
throughout the nose region over time, while in dry machining
notching at the minor cutting edge on the nose region dominated
the wear and its progression.
Fig. 8 shows the crater wear development after 4 min of dry, MQL
and cryogenic machining. Chip-to-tool welding and fracture were
observed on the rake face of the cutting tool used in dry machining.
In MQL machining, fractures on the rake face of the cutting tool
were observed, and this took place at the depth of cut line on the
minor cutting edge. Abrasive marks (grooves) were observed on
the rake face of the cutting tools used in cryogenic machining.
In dry machining, aking (chipping) on the nose region was
observed after 4 min of machining, as shown in Fig. 7. The measured progressive tool-wear with cutting time is given in Fig. 9.
These measured values also conrm the rapid initial tool-wear in
both MQL and dry conditions.

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

99

Fig. 5. Wear on the rake face of the tool after 4 min of machining under dry, MQL, and cryogenic machining of NiTi Alloy I.

3.1.3. Alloy III: VIM + VAR melted, hot-rolled + full anneal


Fig. 10 shows the progressive tool-wear in machining the NiTi
Alloy III under dry, MQL and cryogenic cooling conditions. More
uniform and severe tool-wear and progression was observed in
machining of Alloy III, compared to the other two alloys. While the
wear-rate and progression show a similar tendency in dry and MQL
machining, cryogenic machining made a remarkable contribution
in reducing the progression of tool-wear over the time.
The initial wear in machining of Alloy III was very rapid in dry
and MQL conditions. Severe wear and the resulting change in the
cutting geometry were obvious after only 1 min of cutting time in
both the dry and MQL machining. The measured values of nose
wear of Alloy III with cutting time are presented in Fig. 11. Crater
wear development after 4 min of dry, MQL and cryogenic machining of NiTi Alloy III is shown in Fig. 12. In dry and MQL machining
conditions, major fractures are observed on the rake face and nose
region of the cutting tool that altered the nose geometry. In cryogenic machining, chip-tool welding (chip debris stuck to the cutting
edges) was observed due to the deep grooves on the rake face of
the cutting tool. Grooves on the ank face of cutting tool were also
evident in cryogenic machining.
4. The effect of progressive tool-wear on surface roughness
of machined samples
In biomedical applications, where NiTi alloys are widely used,
high surface quality is one of the signicant expectations and
requirements of the industry. Therefore, it is important to characterize the effects of progressive tool-wear on surface roughness in
machining NiTi alloys. The extremely high tool-wear was resulted
in poor surface topography of the machined workpieces. The difference between the surface topography after the rst and fourth
minutes is large for all the three machining conditions for NiTi Alloy
I as shown in Fig. 13. Similar trends were also observed for Alloys II

and III, hence only the data for Alloy I is presented for brevity. It is
evident that the topography of surface is induced from the progression of tool-wear. During the rst minute of cutting, progression of
tool-wear is quick in dry and MQL machining as compared to cryogenic machining which, leads to an undesired topography of the
machined surface with dry and MQL cutting. Scratches and inconsistent feed marks can be seen in Fig. 13 on the surface of dry and
MQL machined samples after the rst minute of machining time.
Cryogenic machining resulted in smoother and consistent feed
marks on the surface. Abrasive wear with deep grooves was the
major wear mechanism and after 4 min of cutting a large difference between the peak and valley corresponding to an increased
surface roughness was observed on both dry and MQL machining
conditions. Remarkable increases in surface roughness of machined
parts were observed after cutting for 4 min as compared to cutting for 1 min. In the majority of tests, the surface roughness was
more than doubled after cutting for 3 min, indicating that toolwear plays a substantial role on the variation of surface roughness
which is shown in Table 2. Cryogenic machining presents more
promising results in terms of surface quality as compared to dry
and MQL machining at the rst minute of cutting process; however, no remarkable difference between cryogenic and other two
approaches was observed after the fourth minute. A relationship
between the initial conditions of alloys and the machining induced
surface roughness values cannot be established.
5. Analysis and discussion
This current study presents important ndings which will help
to understand and establish the tool-wear mechanisms in machining of NiTi alloys. The progressive tool-wear is analyzed and
discussed considering two main variables: (i) machining conditions
and (ii) the alloy fabrication and processing procedures.
5.1. Machining conditions dry vs. MQL vs. cryogenic
This study conrms the ndings of previous studies presented
by Kaynak et al. (2013b) where cryogenic machining of NiTi alloy is
found to signicantly improve the tool performance in comparison
with dry and MQL machining techniques.

Table 2
Average surface roughness, Ra , of Alloys I, II, and III after 1 and 4 min of cutting time
in dry, MQL, and cryogenic machining. Unit is given in micrometer.
Alloy

Fig. 6. Major cutting edge of cutting tool after 4 min of dry machining of NiTi
Alloy I.

I
II
III

Dry

MQL

Cryogenic

1 min

4 min

1 min

4 min

1 min

4 min

0.17
0.40
0.53

0.62
0.75
0.66

0.29
0.24
0.34

0.70
0.47
0.52

0.13
0.22
0.21

0.52
0.65
0.58

100

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

Fig. 7. Images of wear progression at the nose region in machining of NiTi Alloy II.

Using identical cutting conditions and parameters at 25 m/min


cutting speed, Kaynak et al. (2013b) demonstrated that dry cutting
results in very high workpiece temperatures due to generated heat
while cryogenic machining helps to keep the workpiece temperature low. Specically, the recorded maximum temperature was
523 C, in 259 C and 96 C in dry, MQL and cryogenic machining of room-temperature-martensite NiTi alloy. Therefore, it is not
surprising that the room-temperature austenite NiTi alloys investigated herein exhibited similar tool-wear behavior to that observed

in our prior study, although the room temperature phase state of


NiTi alloy was different.
At temperatures above 200 C (as were observed in both the
dry and MQL machining processes), phase transformation is not
expected to occur in NiTi alloys. Instead, the workpiece will be
in stable austenite phase that behaves like other conventional
elasticplastic metals. Conversely, in the cryogenic machining,
the material exhibits phase-transforming superelastic behavior
that follows the ClausiusClapeyron relationship of increasing

Fig. 8. Crater wear development after 4 min machining under dry, MQL, and cryogenic machining of NiTi Alloy II.

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

Fig. 9. The progression of maximum ank wear at the nose region with time under
dry, MQL, and cryogenic conditions in machining NiTi Alloy II.

transformation stress with increasing temperature. The stress


required to phase-transform the material during cryogenic machining was much less than the yield stress of the stable austenite
during dry and MQL machining. The recorded force data supports this argument. For instance, the average of measured main
cutting force at the rst 10 s during machining of NiTi Alloy I
was 141, 133 and 97 N in dry, MQL, and cryogenic machining,

101

Fig. 11. The progression of maximum ank wear at the nose region with time after
dry, MQL and cryogenic machining of NiTi Alloy III.

respectively. The reduced force requirement to cut this material


with cryogenic machining indicates the role of phase state and
resulting altered friction conditions during the machining process. The elevated stresses during the dry and MQL machining
conditions correlate well with the observations of notch wear
and fracture at the depth of cut line in the nose region, both of
which are indicators of a work hardening phenomenon. Kaynak
et al. (2014b) reported stress-strain response of NiTi (50.8 at

Fig. 10. Optical images of wear progression at the nose region after machining of NiTi Alloy III.

102

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

Fig. 12. Crater wear development after 4 min of dry, MQL and cryogenic machining of NiTi Alloy III.

Fig. 13. Surface topography of machined Alloy I.

%Ni) alloy showing that at low temperature (70 C), yield and ultimate compression stress of NiTi (50.8 at %Ni) are lower than the
ones observed at high temperature (350 C). This is attributed to
difference in deformation mechanism, where primarily twinning
and slip deformation with less stress requirement takes place at
low temperatures in martensite; while slip deformation takes place
at high temperatures in austenite (Kaynak et al., 2014b). According
to Oxley (1962), the large tool-chip contact length is the result of
the high friction. An increase in tool-chip contact length during
dry (380 m) and MQL (314 m) machining compared to cryogenic machining (only 200 m) indicates lower frictional forces

during cryogenic machining (Kaynak et al., 2013b). Our opinion is


that the reduced tool-chip contact length in cryogenic machining
is due to a combination of the lubrication effect of liquid nitrogen (which decreased the coefcient of friction), and the change in
deformation mechanisms of NiTi alloy at cryogenic temperatures
(which decreased the frictional normal force). Since lower frictional
forces and less hardening were expected in the phase-transforming
conditions generated during cryogenic machining, the superior
machining performance was witnessed. These ndings, especially
the improved cutting tool performance with cryogenic machining, are also in good agreement with the studies focused on

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

103

Fig. 14. Comparison of the tool-wear characteristics for the three NiTi Alloys and three machining techniques.

Fig. 15. SEM image of VAR melted and VIM + VAR melted NiTi alloy (NDC marketing, 2012).

machining of other difcult-to-machine materials (Bermingham


et al., 2011).

Two key variables in the alloy manufacturing procedure were


examined in this research: (i) the role of processing, and (ii) the
role of non-metallic inclusions in the microstructure. A comparison
of the tool-wear characteristics amongst the three NiTi Alloys is
presented in Fig. 14, and discussed in the following sections.

conditions were different. Alloy I was cold-worked followed by


a superelastic heat treatment, whereas Alloy II was hot-rolled
followed by full annealing. While variations in their tool-wear performance were observed, those differences were minor. Alloy II
(HR + Full Anneal) exhibited slightly elevated tool-wear compared
to Alloy I in dry and cryogenic machining conditions, while the
opposite is observed in MQL machining. The increased tool-wear
in Alloy II is most likely due to the increased work hardening that
is expected due to the additional ductility of the alloys in the fully
annealed condition.

5.2.1. Role of thermo-mechanical processing NiTi Alloy I vs. II


NiTi Alloys I and II were melted identically and have identical proles of inclusions, but their thermo-mechanical processing

5.2.2. Role of inclusions NiTi Alloy II vs. III


NiTi Alloys II and III had similar thermo-mechanical treatments
(hot-rolled followed by full annealing), but differed in their melting

5.2. Alloy manufacturing procedures

104

Y. Kaynak et al. / Journal of Materials Processing Technology 215 (2015) 95104

technique and therefore inclusion prole. Alloy II was melted by


a VAR process that imparts only oxide inclusions in the material.
Alloy III was melted by VIM + VAR, which resulted in both carbide
and oxide inclusions. Compared to Alloy III, Alloy II has fewer, and,
on average, larger inclusions. While the inclusion area fraction is
generally the same, the inclusion abundance per mm2 in Alloy III
melted by a VIM + VAR process is approximately ve times more
than Alloy I and II melted by a VAR process (NDC Marketing, 2012).
SEM image of VAR melted and VIM + VAR melted NiTi are shown
in Fig. 15.
The tool-wear was markedly increased in Alloy III (VIM + VAR
melted) compared to Alloy II (VAR melted) for all three machining
methods. This trend suggests that the TiC inclusions present in Alloy
III are more detrimental to tool-life than the oxide inclusions. Since
the carbides are harder and generally sharper than oxides, they are
expected to decrease the tool-life, which is in good agreement with
the presented results.
6. Conclusions
In this study, the analysis of new ndings on the effects of
processing and machining conditions on the progressive tool-wear,
and resulting surface quality and phase state of work materials in
machining of room-temperature austenitic NiTi alloys, leads to the
following conclusions:
Alloy considerations
Role of cold work a minor decrease in the machinability of fully
annealed vs. superelastic specimens was observed.
Role of inclusions NiTi alloy containing TiC inclusions resulted in
signicantly decreased tool-life compared to alloys that exclude
these particles. This nding indicates that there is a machinability
benet to selecting carbon-free NiTi alloys.
Machining considerations
Role of cooling/lubrication the new ndings demonstrate that
cryogenic machining is the most promising technique to improve
the machinability of this alloy as it reduces the tool-wear substantially.
Surface roughness due to increased tool-wear over the time,
the surface roughness of machined NiTi alloy deteriorates. With
short machining times, namely 1 min, cryogenic machining has a
superior surface compared to dry and MQL machining. However,
after 4 min, all three machining techniques resulted in nominally
equivalent surface roughness.

Acknowledgments
Authors would like to thank to Nitinol Devices & Components,
Inc. for providing work materials for this study. Support from the

NASA EPSCoR Program under Grant no. NNX11AQ31A is greatly


acknowledged.
References
Bermingham, M., Kirsch, J., Sun, S., Palanisamy, S., Dargusch, M., 2011. New observations on tool life, cutting forces and chip morphology in cryogenic machining
Ti6Al4V. Int. J. Mach. Tool Manuf. 51 (6), 500511.
Dhar, N., Paul, S., Chattopadhyay, A., 2001. The inuence of cryogenic cooling on tool
wear, dimensional accuracy and surface nish in turning AISI 1040 and E4340C
steels. Wear 249 (1011), 932942.
Dhar, N., Paul, S., Chattopadhyay, A., 2002. Machining of AISI 4140 steel under cryogenic cooling tool wear, surface roughness and dimensional deviation. J. Mater.
Process. Technol. 123 (3), 483489.
Duerig, T., Pelton, A., Stockel, D., 1999. An overview of nitinol medical applications.
Mater. Sci. Eng. A 273, 149160.
Kaynak, Y., 2014a. Evaluation of machining performance in cryogenic machining of
Inconel 718 and comparison with dry and MQL machining. Int. J Adv. Manuf.
Technol. 72, 919933.
Kaynak, Y., 2014b. Machining and phase transformation response of roomtemperature austenitic NiTi shape memory alloy. J. Mater. Eng. Perform.,
http://dx.doi.org/10.1007/s11665-014-1058-9.
Kaynak, Y., Karaca, H.E., Noebe, R.D., Jawahir, I.S., 2013a. Analysis of tool-wear and
cutting force components in dry, preheated, and cryogenic machining of NiTi
shape memory alloys. Procedia CIRP 8, 498503.
Kaynak, Y., Karaca, H.E., Noebe, R.D., Jawahir, I.S., 2013b. Tool-wear analysis in
cryogenic machining of NiTi shape memory alloys: a comparison of tool-wear
performance with dry and MQL machining. Wear 306 (12), 5163.
Kaynak, Y., Lu, T., Jawahir, I.S., 2014a. Cryogenic machining-induced surface
integrity: a review and comparison with dry, MQL, and ood-cooled machining.
Mach. Sci. Technol. 18 (2), 149198.
Kaynak, Y., Karaca, H.E., Jawahir, I.S., 2014b. Cutting speed dependent microstructure and transformation behavior of NiTi alloy in dry and cryogenic machining
processes. Metall. Mater. Trans. A (submitted for publication).
Klocke, F., Settineri, L., Lung, D., Claudio Priarone, P., Arft, M., 2013. High performance
cutting of gamma titanium aluminides: inuence of lubricoolant strategy on tool
wear and surface integrity. Wear 302 (12), 11361144.
Kong, M.C., Axinte, D., Voice, W., 2011. Challenges in using waterjet machining of NiTi
shape memory alloys: an analysis of controlled-depth milling. J. Mater. Process.
Technol. 211 (6), 959971.
Kong, M.C., Srinivasu, D., Axinte, D., Voice, W., McGourlay, J., Hon, B., 2013. On
geometrical accuracy and integrity of surfaces in multi-mode abrasive waterjet machining of NiTi shape memory alloys. CIRP Ann. Manuf. Technol. 62 (1),
555558.
Machado, L.G., Savi, M.A., 2003. Medical applications of shape memory alloys. Braz.
J. Med. Biol. Res. 36, 683691.
Morgan, N., 2004. Medical shape memory alloy applications the market and its
products. Mater. Sci. Eng. A 378 (1-2), 1623.
NDC Marketing, 2012. Extra-Low Interstitials Nitinol, www.nitinol.com
Oxley, P., 1962. An analysis for orthogonal cutting with restricted tool-chip contact.
Int. J. Mech. Sci. 4, 129135.
Pusavec, F., Hamdi, H., Kopac, J., Jawahir, I.S., 2011. Surface integrity in cryogenic
machining of nickel based alloy Inconel 718. J. Mater. Process. Technol. 211
(4), 773783.
Saffari, P., Senthilnathan, K., Pfetzing, J., Robertson, S., Pelton, A., 2013. Inuence of
inclusions on localized stress/strain distributions. In: Shape Memory & Superelastic Technologies (SMST) Conference Prague, Czech Republic.
Shahabi, H., Ratnam, M., 2009. Assessment of ank wear and nose radius wear from
workpiece roughness prole in turning operation using machine vision. Int. J.
Adv. Manuf. Technol. 43 (1-2), 1121.
Weinert, K., Petzoldt, V., 2004. Machining of NiTi based shape memory alloys. Mater.
Sci. Eng. A 378 (12), 180184.
Weinert, K., Petzoldt, V., Kotter, D., 2004. Turning and drilling of NiTi shape memory
alloys. CIRP Ann. Manuf. Technol. 53 (1), 6568.

Você também pode gostar