ArticlePDF Available

Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy by Fine Particle Peening

Authors:
  • Toyota National College of Technology

Abstract and Figures

Fine particle peening (FPP) using hydroxyapatite (HAp) shot particles can form a HAp layer on room-temperature substrates by the transfer and microstructural modification of the shot particles. In this study, FPP with HAp shot particles was applied to form a HAp surface layer and improve the fatigue properties of Ti–6Al–4V extra-low interstitial (ELI) for use in bio-implants. The surface microstructures of the FPP-treated specimens were characterized by micro-Vickers hardness testing, scanning electron microscopy, energy-dispersive X-ray spectrometry, X-ray diffraction, and X-ray photoelectron spectroscopy. FPP with HAp shot particles successfully formed a HAp layer on the surface of Ti–6Al–4V ELI in a relatively short period by shot particle transfer at room temperature; however, the thickness and elemental composition of the HAp layer were independent of the FPP treatment time. The original HAp crystal structure remained in the surface-modified layer formed on Ti–6Al–4V ELI after FPP. Furthermore, FPP increased the surface hardness and generated compressive residual stresses at the treated surface of Ti–6Al–4V ELI. Four-point bending fatigue tests were performed at stress ratios of 0.1 and 0.5 to examine the effect of FPP with HAp shot particles on the fatigue properties of Ti–6Al–4V ELI. The fatigue life of the FPP-treated specimen was longer than that of the un-peened specimen because of the formation of a work-hardened layer with compressive residual stress. However, no clear improvement in the fatigue limit of Ti–6Al–4V ELI occurred after FPP with HAp shot particles because of subsurface failures from characteristic facets.
Content may be subject to copyright.
Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy
by Fine Particle Peening
Paper:
Formation of Hydroxyapatite Layer
on Ti–6Al–4V ELI Alloy by Fine Particle Peening
Shoichi Kikuchi1,, Yuki Nakamura2, Koichiro Nambu3, and Toshikazu Akahori4
1Department of Mechanical Engineering, Graduate School of Engineering, Kobe University
1-1 Rokkodai-cho, Nada-ku, Kobe 657-8501, Japan
Corresponding author, E-mail: kikuchi@mech.kobe-u.ac.jp
2Department of Mechanical Engineering, National Institute of Technology, Toyota College, Toyota, Japan
3Toyota Technological Institute, Nagoya, Japan
4Department of Materials Science and Engineering, Faculty of Science and Technology, Meijo University, Nagoya, Japan
[Received January 16, 2017; accepted May 11, 2017]
Fine particle peening (FPP) using hydroxyapatite
(HAp) shot particles can form a HAp layer on
room-temperature substrates by the transfer and mi-
crostructural modification of the shot particles. In
this study, FPP with HAp shot particles was applied
to form a HAp surface layer and improve the fatigue
properties of Ti–6Al–4V extra-low interstitial (ELI)
for use in bio-implants. The surface microstructures
of the FPP-treated specimens were characterized by
micro-Vickers hardness testing, scanning electron mi-
croscopy, energy-dispersive X-ray spectrometry, X-
ray diffraction, and X-ray photoelectron spectroscopy.
FPP with HAp shot particles successfully formed a
HAp layer on the surface of Ti–6Al–4V ELI in a rela-
tively short period by shot particle transfer at room
temperature; however, the thickness and elemental
composition of the HAp layer were independent of the
FPP treatment time. The original HAp crystal struc-
ture remained in the surface-modified layer formed
on Ti–6Al–4V ELI after FPP. Furthermore, FPP in-
creased the surface hardness and generated compres-
sive residual stresses at the treated surface of Ti–6Al–
4V ELI. Four-point bending fatigue tests were per-
formed at stress ratios of 0.1 and 0.5 to examine the
effect of FPP with HAp shot particles on the fatigue
properties of Ti–6Al–4V ELI. The fatigue life of the
FPP-treated specimen was longer than that of the un-
peened specimen because of the formation of a work-
hardened layer with compressive residual stress. How-
ever, no clear improvement in the fatigue limit of Ti–
6Al–4V ELI occurred after FPP with HAp shot parti-
cles because of subsurface failures from characteristic
facets.
Keywords: fine particle peening, titanium alloy, hydrox-
yapatite, fatigue, biomaterial
1. Introduction
The Ti–6Al–4V alloy is widely used in various engi-
neering fields because it possesses a high specific strength
and good heat resistance. In particular, Ti–6Al–4V extra-
low interstitial (ELI) is used as a substitute for hard bi-
ological tissues in biomaterials such as artificial joints,
dental implants, and fracture fixators, because it exhibits
excellent corrosion resistance, high tissue compatibility,
and a lower Young’s modulus than ferrous materials [1].
Osteoconductivity, or the characteristics permitting bone
growth on a material surface, is required to affix titanium-
based bio-implants [2] to human bones over a long period
of time.
Surface modification processes without the use of bone
cement have been introduced to improve the osteocon-
ductivity and bonding strength between human bones
and bio-implants [3–12] because the surfaces of bio-
implants are in contact with body tissues. For exam-
ple, Kokubo et al. [3, 4] investigated the effects of pre-
treatment with alkali hydroxide solutions on the forma-
tion of hydroxyapatite (HAp) on commercially pure (CP)
titanium in simulated body fluid to produce a bioactive
material surface. Plasma-sprayed HAp coatings are ef-
fective for improving the hard tissue compatibility of
titanium-based bio-implants because they form thick HAp
layers [5–11]. However, the thermal energy induced by
high-temperature and energetic processes, such as plasma
spraying, can change the crystalline structure of HAp [10]
or the titanium substrate [13], which negatively affects the
HAp-coated bio-implants.
To address this problem, various room-temperature
HAp coating methods have been developed for titanium-
based bio-implants [14–20]. Ishikawa et al. [17, 18] re-
ported the homogeneous surface coating of a titanium
plate with HAp using an ordinary sandblaster at room
temperature. In our previous study [20], fine particle
peening (FPP) with HAp shot particles was used to form
a HAp layer on a CP titanium plate at room temperature
through shot particle transfer [21–25] induced by the high
particle velocities used for FPP [26, 27]. Another impor-
tant aspect of HAp coatings formed on titanium-based
bio-implants is their behavior under cyclic loading, like
that applied during use in the human body. Kangasniemi
et al. [11] showed that fracture occurred at the interface
between the coating and titanium substrate by investigat-
Int. J. of Automation TechnologyVol.11 No.6, 2017 915
Kikuchi, S. et al.
5Pm
5 ȝm
Fig. 1. Image quality (IQ) map obtained from EBSD analy-
sis of Ti–6Al–4V ELI alloy.
ing the mechanical properties of bioactive coating mate-
rials using a developed testing system. In contrast, the
HAp layer remained on the FPP-treated CP titanium af-
ter fatigue testing without delamination on the fracture
surface [20]. Furthermore, FPP can improve the fatigue
properties of the titanium substrate by increasing the sur-
face hardness and forming fine grains [28–30].
In the present study, FPP using HAp shot particles was
used to form a HAp surface layer and improve the fatigue
properties of Ti–6Al–4V ELI, which is often applied in
bio-implants. The purpose of this study was to charac-
terize the HAp layer formed on Ti–6Al–4V ELI by FPP
using HAp shot particles at room temperature and to ex-
perimentally examine the fatigue properties of the resul-
tant coated alloy under four-point bending.
2. Experimental Procedures
2.1. Material and Specimen Preparation
The material used in this work was Ti–6Al–4V (ELI
grade) with the chemical composition of 6.31% alu-
minum, 4.13% vanadium, 0.12% iron, 0.002% hydrogen,
0.006% nitrogen, 0.11% oxygen, and 0.024% carbon by
mass, with titanium comprising the balance. Fig. 1 shows
an image quality (IQ) map obtained by electron backscat-
ter diffraction (EBSD) analysis for Ti–6Al–4V ELI con-
taining both the equiaxed
α
-phase and
β
-phase. This ma-
terial has the Vickers hardness of 340.2 ±4.7 HV, as mea-
sured for a polished surface with an indentation force of
0.098 N and a load holding time of 5 s (n=30). Ti–6Al–
4V ELI plates of 11 mm in thickness were machined into
1.5-mm-thick sheets and then cut into 3 ×20 mm spec-
imens using a wire electrical discharge machine. After
machining, the specimens were polished with emery pa-
per (#320 to #4000) to 1 mm in thickness and then to a
mirror finish using a SiO2suspension. The sides of the
specimen were also polished with emery paper (#500) to
remove the electro-discharge machined layer.
Table 1. Conditions for FPP.
Peening pressure 0.6 MPa
Peening time 1, 10, 20, 30 s
Nozzle distance 50 mm
Table 2. Residual stress measurement conditions.
Tube voltage 40 kV
Tube current 30 mA
Diffraction angle 2
θ
154.3 deg.
Diffraction plane (331)
Incident angle 10, 20, 30, 35, 40 deg.
Beam diameter 1 mm
Stress constant 170.77 MPa/deg.
FPP was performed on the polished specimens using a
direct pressure-type apparatus under the conditions given
in Tabl e 1 at room temperature in air. The shot particles
with diameters of 50
μ
m [20] were produced by pulveriz-
ing HAp (Ca10(PO4)6(OH)2) fabricated by ECCERA Co.,
Ltd. After performing FPP, the specimens were placed in
an ultrasonic bath of acetone for 600 s to remove free par-
ticles from the surfaces.
2.2. Characterization of the Surface-Modified
Layer
The surface microstructures of the specimens were
characterized using scanning electron microscopy (SEM)
with an accelerating voltage of 15 kV. The FPP-treated
surfaces were also analyzed using energy dispersive X-ray
spectrometry (EDX) in an area of 1.13 mm2observed at
100×magnification with an accelerating voltage of 20 kV,
as well as X-ray photoelectron spectroscopy (XPS) with
Mg K
α
radiation. The crystal structures of the speci-
mens were identified using X-ray diffraction (XRD) with
Cu K
α
radiation. The HAp layer was also analyzed us-
ing EDX at 5000×magnification in longitudinal cross-
sections.
The hardness distributions were measured along longi-
tudinal cross-sections of the FPP-treated specimens using
a micro-Vickers hardness tester with an indentation force
of 0.098 N and a load holding time of 10 s. The residual
stress was also measured at the top surface of a transverse
section of the specimen using XRD with Co K
α
radia-
tion and a position-sensitive proportional counter (PSPC)
system based on the sin2
ψ
method (n=2) [31, 32]. The
conditions for the residual stress measurement are shown
in Tabl e 2 .
2.3. Fatigue Tests Under Four-Point Bending
Fatigue tests were conducted in an electrodynamic fa-
tigue testing machine (loading capacity: 500 N) under
four-point bending using the mirror-finished and FPP-
treated specimens measuring 3×20×1 mm at the stress
916 Int. J. of Automation TechnologyVol.11 No.6, 2017
Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy
by Fine Particle Peening
(a) 0 s (Un-peened)
(b) 1 s (c) 10 s
(d) 20 s (e) 30 s
50 μm
50 μm50 μm
50 μm50 μm
Fig. 2. SEM micrographs of (a) un-peened specimen and
specimens treated with FPP for (b) 1 s, (c) 10 s, (d) 20 s, and
(e) 30 s.
ratios of 0.1 and 0.5, because fatigue crack propagation
in titanium alloys is influenced by the stress ratio [33–
37]. The frequency of stress cycling was 10 Hz based on
the JIS T 0309 standard and the tests were conducted in
air without temperature or moisture control. The fatigue
limit was defined based on the JSMS standard [38]. After
testing, the fracture surfaces of the failed specimens were
observed using SEM at an accelerating voltage of 10 kV.
3. Results and Discussion
3.1. Transfer of HAp Shot Particles by FPP to
Ti–6Al–4V ELI
Figure 2 shows backscattered electron (BSE) micro-
graphs of the un-peened and FPP-treated specimen sur-
faces, in which the contrast is related to the composi-
tion of the surface layer. The bright regions correspond
to the titanium substrate because the brightness increases
with increasing atomic number. A smooth surface with
high brightness is observed for the un-peened specimen,
as shown in Fig. 2(a), whereas the contrast in the BSE
images for the FPP-treated specimens is clearly different
(Figs. 2(b)–(e)). This is attributed to the presence of ir-
regularly shaped HAp shot particles transferred onto the
FPP-treated surface.
Figure 3 shows EDX maps of an un-peened specimen
surface and the surface of a specimen treated with FPP
for 30 s after ultrasonic cleaning. Only the substrate ele-
ments are detected in the un-peened specimen (Fig. 3(a)).
In contrast, calcium, phosphorus, and oxygen, present in
CaPO
(a) Un-peened
Ti
(b) FPP (30 s)
Not detected
Not detected
Not detected
200 μm
Fig. 3. EDX maps for (a) un-peened and (b) FPP-treated
specimen surfaces (peening time: 30 s).
FPP treatment time, s
0
100
Composition, mass%
20
40
60
80
0102030
P
Ca
OTi
Al
V
Fig. 4. Relationship between peening time and elemental
composition of FPP-treated surfaces determined from EDX
analysis.
the HAp shot particles, are also detected on the specimen
surfaces treated with FPP for 30 s (Fig. 3(b)). It is consid-
ered that the layer with these elements corresponds to the
transferred HAp layer. In addition, a trace amount of ti-
tanium is detected on the FPP-treated surface because the
elemental composition at depths of several micrometers
was also detected in the EDX analysis; however, the HAp
layer remained after performing ultrasonic cleaning.
To investigate the formation of the HAp layer dur-
ing FPP, specimens treated with FPP for various peening
times were analyzed using EDX. Fig. 4 shows the rela-
Int. J. of Automation TechnologyVol.11 No.6, 2017 917
Kikuchi, S. et al.

(Surface)
5 μm
CaPOTi
Not detected
Not detected
Not detected
(a) 0 s (Un-peened)
FPP treatment time
(b) 1 s (c) 10 s (d) 20 s (e) 30 s
Fig. 5. Longitudinal EDX maps for (a) un-peened and (b)–(e) FPP-treated specimens.
tionship between the elemental composition of the surface
layer analyzed by EDX and the FPP treatment time. The
data for an FPP treatment time of 0 s is similar to that
of the un-peened specimen shown in Fig. 3(a). Calcium,
phosphorus, and oxygen are detected in each FPP-treated
specimen, although there is no noticeable dependence of
the content of each element on the FPP treatment time.
This lack of influence of the FPP treatment time on the
content of each element in the HAp layer was also ob-
served in CP titanium [20]. Fig. 5 shows longitudinal
EDX maps of the un-peened and FPP-treated specimens.
For the un-peened specimen (Fig. 5(a)), only the substrate
elements are detected, which is consistent with the results
shown in Figs. 3(a) and 4. In contrast, calcium, phospho-
rus, and oxygen from the HAp shot particles are detected
near the surfaces treated with FPP, while titanium is not
detected (Figs. 5(b)–(e)).
Figure 5 also demonstrates that the HAp layer is non-
homogeneously formed at the FPP-treated surface. To ex-
amine the effect of the FPP treatment time on the thick-
ness of the HAp layer formed on Ti–6Al–4V quantita-
tively, the equivalent thickness of the HAp layer was cal-
culated from Eq. (1) if the HAp layer, consisting of the
HAp shot particles transferred onto the surface, was uni-
formly formed over the entire surface:
teq =A
b,............... (1)
where teq is the equivalent thickness of the HAp layer
[
μ
m], Ais the area of detection for the calcium element
[
μ
m2], and bis the width of the specimen in the analyzed

FPP treatment time, s
10
Thickness of the HAp layer, Pm
0102030
0
2
4
6
8
Un-peened
FPP
(0.04 Pm/s)
(6.90 Pm/s)
Fig. 6. Relationship between peening time and equivalent
thickness of the HAp layer, estimated by Eq. (1).
area (25.2
μ
m).
Figure 6 shows the relationship between the equivalent
thickness of the HAp layer estimated by Eq. (1) and the
FPP treatment time. The HAp layer thickness increases
significantly for FPP treatment times increasing to 1 s,
withtheincreaserateof6.90
μ
m/s. This result indicates
that FPP can form a HAp layer on Ti–6Al–4V ELI in a
short period at room temperature. In contrast, there is no
noticeable difference in the HAp layer thickness for FPP
treatment times greater than 1 s; the thickness increases
only slightly at the rate of 0.04
μ
m/s to the HAp layer
thickness of approximately 7
μ
m. Thus, the EDX analy-
ses suggest that FPP can form a HAp layer in a relatively
918 Int. J. of Automation TechnologyVol.11 No.6, 2017
Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy
by Fine Particle Peening
Un-peened
FPP (20 s)
Binding energy, eV
452467 462 457
Counts
Ti 2p
3/2
(TiO
2
)
O 1s (Phosphate)
Binding energy, eV
525540 535 530
Counts
FPP (20 s)
(a) Ti 2p (b) O 1s
Un-peened
O 1s (Ti-O)
O 1s (OH)
Mg KDMg KD
Ti 2p
1/2
(TiO
2
)
Fig. 7. XPS (a) Ti 2pand (b) O 1sspectra for un-peened
and FPP-treated specimens (peening time: 20 s).
short period (1 s) at room temperature, and that the thick-
ness and elemental composition of the HAp layer are in-
dependent of the FPP treatment time.
3.2. Characterization of HAp Layer Formed by
FPP on Ti–6Al–4V ELI
XPS and XRD analyses were conducted to examine the
microstructures of the surface HAp layers in more de-
tail. Fig. 7 shows the XPS Ti 2pand O 1sspectra for
the un-peened specimen and the specimen treated with
FPP for 20 s. No Ti 2 ppeaks are detected for the FPP-
treated specimen, whereas Ti 2ppeaks (TiO2) are clearly
evident for the un-peened specimen (Fig. 7(a)). Further-
more, an O 1speak (Ti–O) is detected from the un-peened
specimen, corresponding to the Ti 2 ppeak, as shown in
Fig. 7(b). In contrast, no titanium oxide peaks are de-
tected for the FPP-treated specimen, whereas O 1s(phos-
phate) peaks are clearly evident because of the formation
of the HAp layer at the top surface. These results indicate
that the HAp layer is formed over the entire surface of the
FPP-treated specimen.
Figure 8 shows the XRD patterns for an un-peened
specimen and a specimen treated with FPP for 20 s.
The un-peened specimen has diffraction peaks from the
substrate corresponding to the
α
-Ti and
β
-Ti phases
(Fig. 8(a)). However, XRD peaks associated with
HAp are also detected from the FPP-treated specimen
(Fig. 8(b)). The intensity of both the
α
-Ti and
β
-Ti peaks
observed from the FPP-treated specimen is decreased
slightly following FPP, and the Ti peaks are shifted to
lower angles. These changes are attributed to the plas-
tic deformation of the substrate during FPP, as discussed
in the next section.
The XPS and XRD analyses suggest that the origi-
nal HAp crystal structure remains in the surface-modified
layer formed on Ti–6Al–4V ELI after performing FPP.
3.3. Hardness and Residual Stress Measurements
of FPP-treated Ti–6Al–4V ELI
Figure 9 shows the distribution of Vickers hardness at
various depths in longitudinal sections of the FPP-treated
specimens. The specimen treated for 1 s exhibits al-
most the same hardness as the substrate (340.2 ±4.7 HV),
Cu KD
30 35 45
Intensity
Diffraction angle 2
T
, degree
40
E-Ti HAp
(a) Un-peened
(b) FPP (20 s)
D-Ti
(0002)
(1010)
(1011)
(110)
(202)
(300)
(211)
Fig. 8. XRD patterns for (a) un-peened and (b) FPP-treated
specimens (peening time: 20 s).
whereas the surface hardness values of the specimens
treated for longer than 10 s are higher than that of the sub-
strate. The surface hardness of the FPP-treated specimens
tends to increase with the FPP treatment time up to 20 s
and then remains constant. Furthermore, the thickness
of the surface hardened layer formed in the FPP-treated
specimens tends to increase with the FPP treatment time
up to 20 s.
Figure 10 shows the residual stress generated at the
surfaces treated with FPP for various peening times and
that for the un-peened specimen, as determined by XRD.
In the case of the un-peened specimen, tensile residual
stress is generated, whereas compressive residual stress
is generated at the surface of each FPP-treated speci-
men. This suggests that FPP using HAp shot particles can
generate compressive residual stresses at the surface of
Ti–6Al–4V ELI. In the FPP-treated specimens, the com-
pressive residual stress tends to decrease with increasing
FPP treatment time up to 20 s before increasing again.
The compressive residual stress generally increases as the
thickness of the surface hardened layer decreases; how-
ever, the compressive residual stress increases again for
the specimen treated by FPP for 30 s, although the thick-
ness of the surface hardened layer is almost the same, as
shown in Fig. 9. The same tendency was also observed for
a CP titanium specimen treated with FPP using HAp shot
particles [20] because the FPP treatment time affects both
the residual stress generated on the top surface and the
thickness of the compressive residual stress layer. Fur-
thermore, for excessive FPP treatment times, the degree
of plastic deformation in the substrate increases, and abra-
sion of the substrate and the transfer of HAp shot particles
simultaneously occur at the specimen surface during FPP;
therefore, compressive residual stress are not expected to
change linearly with the FPP treatment time.
Thus, FPP using HAp shot particles can form a work-
hardened layer with a compressive residual stress in Ti–
6Al–4V ELI at varied FPP treatment times.
Int. J. of Automation TechnologyVol.11 No.6, 2017 919
Kikuchi, S. et al.
300
400
320
Vickers hardness, HV (0.098 N)
0
Distance from surface, Pm
50 100 150 200
340
360
380
1 s
20 s
FPP treatment time
30 s
10 s
Un-peened specimen
(340.24.7 HV)
Fig. 9. Distributions of Vickers hardness at various
longitudinal-section depths.
FPP treatment time, s
-500
200
Residual stress, MPa
0 102030
-100
-200
-300
-400
Co KD
100
0
Un-peened
FPP
(Average Standard error)
Fig. 10. Residual stress measured on the treated surface as
a function of peening time.
3.4. Fatigue Properties of FPP-Treated Ti–6Al–4V
ELI Under Four-Point Bending
Four-point bending fatigue tests were conducted for the
specimen treated with FPP for 20 s, which had the lowest
compressive residual stress, as shown in Fig. 10.There-
sults of the fatigue tests under four-point bending are plot-
ted as SNdiagrams for the un-peened and FPP-treated
specimens tested at stress ratios of 0.1 and 0.5 (Fig. 11).
The plots with arrows indicate a run-out specimen with-
out failure. In every specimen series, a fatigue limit is
clearly observed, so that the SNcurve is determined by
using the bilinear SNmodel with fatigue limiting in the
JSMS standard regression models [38]. FPP using HAp
shot particles prolongs the fatigue life of Ti–6Al–4V ELI
compared to that of the un-peened specimen at a stress
ratio of 0.1, because FPP increases the surface hardness
of the Ti–6Al–4V ELI alloy and generates compressive
residual stress, as shown in Figs. 9 and 10.However,no
clear improvement occurs in the fatigue limit of Ti–6Al–
4V ELI by FPP using HAp shot particles. This may be
attributed to the FPP-induced surface roughness, because
the surface roughness affects the mechanical properties
of a specimen [39–41]. The FPP-treated specimen fails at
σ
a=314.3 MPa, whereas fatigue failure does not occur at


200
100
Number of cycles to failure N
f
, cycle
Stress amplitude
V
a
, MPa
10
4
10
5
10
6
10
7
10
8
300
400
: Run-out
Four-point bending
*: Subsurface failure
10 Hz
500
*
Un-peened
FPP (20 s)
R= 0.1 R= 0 .5
Fig. 11. Four-point bending fatigue test results for un-
peened and FPP-treated specimens in terms of stress ampli-
tude (peening time: 20 s).


700
600
Number of cycles to failure Nf, cycle
Maximum stress
V
max, MPa
104105106107108
1100
800
900
1000
Un-peened
FPP (20 s)
R= 0.1 R= 0.5
: Run-out
Four-point bending
*: Subsurface failure
10 Hz
*
Fig. 12. Four-point bending fatigue test results for un-
peened and FPP-treated specimens in terms of maximum
stress (peening time: 20 s).
σ
a=313.8 MPa in the un-peened specimen. When tested
at a stress ratio of 0.5, the FPP-treated specimens show
almost the same fatigue lifetimes and fatigue limits as the
un-peened specimen.
Figure 11 also reveals that the fatigue limit for the
specimens tested at R=0.1 is higher than that for the
specimens tested at R=0.5. This difference in the fa-
tigue limit is attributed to the value of the mean applied
stress. To investigate the effect of the stress ratio on the
fatigue properties of Ti–6Al–4V ELI, the fatigue test re-
sults are replotted as a function of the maximum stress
σ
max as shown in Fig. 12. For stress ratios of 0.1 and 0.5,
the plots of each specimen series are almost within the
same bands. This result implies that the effect of stress
ratio on the fatigue properties of Ti–6Al–4V ELI disap-
pears. A modified Goodman diagram and the Smith–
Watson–Topper (SWT) model [42] were used to exam-
ine the stress ratio effect; however, the fatigue properties
of Ti–6Al–4V ELI were determined using the maximum
stress applied to the specimen’s surface under four-point
bending. Therefore, fitting was performed using the data
for all the un-peened and FPP-treated specimens to statis-
tically investigate the effect of FPP with HAp shot parti-
920 Int. J. of Automation TechnologyVol.11 No.6, 2017
Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy
by Fine Particle Peening
(a) 35 mag.(b) 500 mag.(c) 3000 mag.
Fracture surface 1 Fracture surface 2
Facet Facet
5 μm
Facet
Facet
500 μm
(Tensile side)
Crack initiation site
Crack propagation region
Instantaneous fracture region
(Tensile side)
Crack initiation site
Instantaneous fracture region
Crack propagation region
50 μm
500 μm
50 μm
5 μm
Fig. 13. Typical features of both fracture surfaces of FPP-
treated specimen failing at Nf=4.35×106(stress amplitude:
314.3 MPa, stress ratio: 0.1, peening time: 20 s).
cles on the fatigue properties. Regression SNcurves for
the un-peened and FPP-treated specimens are respectively
expressed by the following formulae:
Un peened :
σ
max =108.7log(N)+1433.7,(2)
FPP treated :
σ
max =122.2log(N)+1543.0,(3)
where
σ
max is the maximum stress applied to the speci-
men surface [MPa], and Nis the number of cycles.
The fatigue life of the FPP-treated specimens was
longer than that of the un-peened specimens. The crit-
ical number of stress cycles giving the fatigue limit for
the FPP-treated specimens (Nw=8.82 ×106) was higher
than that for the un-peened specimens (Nw=5.80 ×106).
However, the fatigue limit of the FPP-treated specimen
(
σ
w,max =694.1 MPa) was almost equal to that of the un-
peened specimen (
σ
w,max =698.7 MPa). Thus, FPP with
HAp shot particles can prolong the fatigue life of Ti–6Al–
4V ELI, but does not improve the fatigue limit. In CP
titanium, FPP with HAp shot particles was reported to in-
crease the fatigue limit [20]; therefore, improvement of
the fatigue limit of titanium by FPP using HAp shot par-
ticles is different.
The fracture surfaces of the failed FPP-treated speci-
mens were observed using SEM to examine the fracture
mechanism. Fig. 13 shows the typical features of both
fracture surfaces observed at various magnifications for
the FPP-treated specimen that failed with a long lifetime
of Nf=4.35 ×106.InFig. 13, the tensile stress has been
applied to the lower surface. Microscopic observation re-
veals that only one fatigue crack is present near the spec-
imen surface; this propagates gradually across the cross-
10
Height, Pm
0
Distance from surface, Pm
51015
8
6
4
2
0
Analyzed line
(Surface)
2 μm
30.8o
Facet
Fig. 14. Three-dimensional analysis of the fracture surface
1 of FPP-treated specimen failed at Nf=4.35 ×106(stress
amplitude: 314.3 MPa, stress ratio: 0.1, peening time: 20 s).
section of the specimen (Fig. 13(a)). The fracture surface
is divided into two regions by a clear boundary, as indi-
cated by the dotted line. In addition, a characteristic facet
is clearly observed at the crack initiation site in the higher-
magnification SEM micrographs (Figs. 13(b) and 13(c)).
The FPP-treated specimen fails from the facet at a stress
ratio of 0.1; crack initiation can be observed below the
FPP-treated surface because the resistance to crack initia-
tion is higher at the high-hardness surface, which reduces
the fatigue limit of the Ti–6Al–4V ELI. In Figs. 11 and
12, the asterisk symbol “*” indicates that the specimen
failed in the subsurface fracture mode.
To examine the fatigue fracture mechanism of the
FPP-treated specimen in more detail, three-dimensional
fracture surfaces were produced by exclusive software
(Alicona’s MeX) for 3D-fracture surface reconstruction.
Fig. 14 shows the profile curve for the fracture surface 1
of the FPP-treated specimen shown in Fig. 13.There-
lationship between the height based on the lowest point
in the analyzed area and the distance from the specimen’s
surface is shown. The inclination angle of the facet ob-
served on the fracture surface is 30.8. The size of the
facet is almost equal to that of the
α
-grain shown in
Fig. 1; therefore, subsurface fracture occurs from a coarse
α
-grain with a weak microstructural orientation. In ac-
cordance with previous investigations on Ti–6Al–4V [43,
44], microstructural inhomogeneities in the phase distri-
bution are responsible for internal crack initiation, which
corresponds to the specific local conditions of plastic de-
formation in the
α
-phase and
β
-phase.
Int. J. of Automation TechnologyVol.11 No.6, 2017 921
Kikuchi, S. et al.
4. Conclusions
FPP was performed to form a HAp shot particle layer
on Ti–6Al–4V ELI at room temperature. The surface mi-
crostructure of the FPP-treated specimen was character-
ized, and the effect of FPP on the fatigue properties of
Ti–6Al–4V ELI was examined under four-point bending
at various stress ratios. The main conclusions of this study
are as follows:
1. FPP using HAp shot particles can form a surface-
modified layer with the original HAp crystal struc-
ture on Ti–6Al–4V ELI in a relatively short period at
room temperature.
2. The thickness and elemental composition of the HAp
layer are independent of the FPP treatment time.
3. FPP with HAp shot particles can increase the fatigue
life of Ti–6Al–4V ELI by forming a work-hardened
layer with compressiveresidual stress on Ti–6Al–4V
ELI.
4. The fatigue limit of Ti–6Al–4V ELI does not in-
crease by FPP using HAp shot particles, because the
FPP-treated specimen fails via subsurface fracture
from the characteristic facet, beginning at a coarse
α
-grain with a weak microstructural orientation. The
fatigue crack then propagates gradually across the
cross-section of Ti–6Al–4V ELI.
Acknowledgements
The authors would like to thank The Inamori Foundation and
the Advanced Machining Technology & Development Associa-
tion for their financial support. FPP using HAp shot particles was
based on Japanese Patent No.3314070, filed by FUJI KIHAN Co.,
Ltd and the Aichi Center for Industry and Science Technology.
References:
[1] M. Long and H. J. Rack, “Titanium Alloys in Total Joint
Replacement-A Materials Science Perspective,” Biomaterials,
Vol.19, pp. 1621-1639, 1998.
[2] A. Rouzrokh, C. Y. H. Wei, K. Erkorkmaz, and R. M. Pillar, “Ma-
chining Porous Calcium Plyphosphate Implants for Tissue Engi-
neering Applications,” Int. J. of Automation Technology, Vol.4,
No.3, pp. 291-302, 2010.
[3] T. Kokubo, F. Miyaji, H. M. Kim, and T. Nakamura, “Spontaneous
Formation of BoneLike Apatite Layer on Chemically Treated Ti-
tanium Metals,” J. of American Ceramics Society, Vol.79, No.4,
pp. 1127-1129, 1996.
[4] H. M. Kim, T. Himeno, M. Kawashita, J. H. Lee, T. Kokubo, and T.
Nakamura, “Surface Potential Change in Bioactive Titanium Metal
during the Process of Apatite Formation in Simulated Body Fluid,”
J. of Biomedical Materials Research Part A, Vol.67, pp. 1305-1309,
2003.
[5] R. B. Heimann, “Thermal Spraying of Biomaterials,” Surface and
Coatings Technology, Vol.201, pp. 2012-2019, 2006.
[6] S. Lazarinis, J. Karrholm, and N. P. Hailer, “Effects of Hydroxya-
patite Coating of Cups Used in Hip Revision Arthroplasty,” Acta
Orthopaedica, Vol.83, pp. 427-435, 2012.
[7] T. Laonapakul, A. R. Nimkerdphol, Y. Otsuka, and Y. Mutoh, “Fail-
ure Behavior of Plasma-Sprayed HAp Coating on Commercially
Pure Titanium Substrate in Simulated Body Fluid (SBF) under
Bending Load,” J. of Mechanical Behavior of Biomedical Mate-
rials, Vol.15, pp. 153-166, 2012.
[8] T. Laonapakul, Y. Otsuka, A. R. Nimkerdphol, and Y. Mu-
toh, “Acoustic Emission and Fatigue Damage Induced in Plasma-
Sprayed Hydroxyapatite Coating Layers,” J. of Mechanical Behav-
ior of Biomedical Materials, Vol.15, pp. 123-133, 2012.
[9] M. Inagaki, Y. Yokogawa, and T. Kameyama, “Bond Strength Im-
provement of Hydroxyapatite/Titanium Composite Coating by Par-
tial Nitriding during RF-Thermal Plasma Spraying,” Surface and
Coatings Technology, Vol.173, pp. 1-8, 2003.
[10] R. L. Reis and F. J. Monteriro, “Crystallinity and Structural Changes
in HA Plasma-Sprayed Coatings Induced by Cyclic Loading in
Physiological Media,” J. of Materials Science and Materials in
Medicine, Vol.7, pp. 407-411, 1996.
[11] I. M. O. Kangasniemi, C. C. P. M. Verheyen, E. A. V. der Velde, and
K. De Groot, “In Vivo Tensile Testing of Fluorapatite and Hydroxy-
lapatite Plasma-Sprayed Coatings,” J. of Biomedical Materials Re-
search Part A, Vol.28, No.5, pp. 563-572, 1994.
[12] M. Mizutani, R. Honda, Y. Kurashina, J. Komotori, and H. Ohmori,
“Improved Cytocompatibility of Nanosecond-Pulsed Laser-Treated
Commercially Pure Ti Surfaces,” Int. J. of Automation Technology,
Vol.8, No.1, pp. 102-109, 2014.
[13] E. Gariboldi and B. Previtali, “High Tolerance Plasma Arc Cutting
of Commercially Pure Titanium,” J. of Materials Processing Tech-
nology, Vol.160, pp. 77-89, 2005.
[14] R. Akatsuka, H. Ishihata, M. Noji, K. Matsumura, T. Kuriyagawa,
and K. Sasaki, “Effect of Hydroxyapatite Film Formed by Powder
Jet Deposition on Dentin Permeability,” European J. of Oral Sci-
ences, Vol.120, pp. 558-562, 2012.
[15] R. Akatsuka, K. Sasaki, M. S. S. Zahmaty, M. Noji, T. Anada,
O. Suzuki, and T. Kuriyagawa, “Characteristics of Hydroxyap-
atite Film Formed on Human Enamel with the Powder Jet Depo-
sition Technique,” J. of Biomedical Materials Research B, Vol.98,
pp. 210-216, 2011.
[16] N. Hayashi, S. Ueno, S. V. Komarov, E. Kasai, and T. Oki, “Fab-
rication of Hydroxyapatite Coatings by the Ball Impact Process,”
Surface and Coatings Technology, Vol.206, pp. 3949-3954, 2012.
[17] T. Mano, Y. Ueyama, K. Ishikawa, T. Matsumura, and K. Suzuki,
“Initial Tissue Response to a Titanium Implant Coated with Apatite
at Room Temperature Using a Blast Coating Method,” Biomateri-
als, Vol.23, pp. 1931-1936, 2002.
[18] K. Ishikawa, Y. Miyamoto, M. Nagayama, and K. Asaoka, “Blast
Coating Method: New Method of Coating Titanium Surface with
Hydroxyapatite at Room Temperature,” J. of Biomedical Materials
Research, Vol.38, pp. 129-134, 1997.
[19] L. O’Neill, C. O’Sullivan, P. O’Hare, L. Sexton, F. Keady, and J.
O’Donoghue, “Deposition of Substituted Apatites onto Titanium
Surfaces Using a Novel Blasting Process,” Surface and Coatings
Technology, Vol.204, pp. 484-488, 2009.
[20] S. Kikuchi, S. Yoshida, Y. Nakamura, K. Nambu, and T. Akahori,
“Characterization of the Hydroxyapatite Layer Formed by Fine Hy-
droxyapatite Particle Peening and its Effect on the Fatigue Prop-
erties of Commercially Pure Titanium under Four-Point Bending,”
Surface and Coatings Technology, Vol.288, pp. 196-202, 2016.
[21] S. Ota, H. Akebono, S. Kikuchi, K. Murai, J. Komotori, K.
Fukazawa, Y. Misaka, and K. Kawasaki, “Surface Modification of
Carbon Steel by Atmospheric-Controlled IH-FPP Treatment Us-
ing Mixed Chromium and High-Speed Steel Particles,” Materials
Transactions, Vol.57, No.10, pp. 1801-1806, 2016.
[22] S. Kikuchi, T. Fukuoka, T. Sasaki, J. Komotori, K. Fukazawa, Y.
Misaka, and K. Kawasaki, “Increasing Surface Hardness of AISI
1045 Steel by AIH-FPP / Plasma Nitriding Treatment,” Materials
Transactions, Vol.54, No.1, pp. 344-349, 2013.
[23] Y. Kameyama and J. Komotori, “Effect of Micro Ploughing dur-
ing Fine Particle Peening Process on the Microstructure of Metallic
Materials,” J. of Materials Processing Technology, Vol.209, No.20,
pp. 6146-6155, 2009.
[24] Y. Kameyama, K. Nishimura, H. Sato, and R. Shimpo, “Effect of
Fine Particle Peening Using Carbon-Black/Steel Hybridized Parti-
cles on Tribological Properties of Stainless Steel,” Tribology Int.,
Vol.78, pp. 115-124, 2014.
[25] Y. Kameyama, H. Ohmori, H. Kasuga, and T. Kato, “Fabrication
of Micro-Textured and Plateau-Processed Functional Surface by
Angled Fine Particle Peening Followed by Precision Grinding,”
CIRP Annals-Manufacturing Technology, Vol.64, No.1, pp. 549-
552, 2015.
[26] T. Ito, S. Kikuchi, Y. Hirota, A. Sasago, and J. Komotori, “Analysis
of Pneumatic Fine Particle Peening Process by Using a High-Speed-
Camera,” Int. J. of Modern Physics B, Vol.24, No.15-16, pp. 3047-
3052, 2010.
[27] S. Kikuchi, Y. Nakamura, K. Nambu, and M. Ando, “Effect of Shot
Peening Using Ultra-Fine Particles on Fatigue Properties of 5056
Aluminum Alloy under Rotating Bending,” Materials Science and
Engineering A, Vol.652, pp. 279-286, 2016.
[28] T. Morita, H. Nakaguchi, S. Noda, and C. Kagaya, “Effects of
Fine Particle Bombarding on Surface Characteristics and Fatigue
Strength of Commercial Pure Titanium,” Materials Transactions,
Vol.53, No.11, pp. 1938-1945, 2012.
922 Int. J. of Automation TechnologyVol.11 No.6, 2017
Formation of Hydroxyapatite Layer on Ti–6Al–4V ELI Alloy
by Fine Particle Peening
[29] T. Morita, K. Asakura, and C. Kagaya, “Effect of Combination
Treatment on Wear Resistance and Strength of Ti-6Al-4V alloy,”
Materials Science and Engineering A, Vol.618, pp. 438-446, 2014.
[30] S. Kikuchi and J. Komotori, “Effect of Fine Particle Peening on At-
mospheric Oxidation Behavior of Ti-6Al-4V Alloy, J. of the Japan
Institute of Metals and Materials, Vol.80, No.2, pp. 114-120, 2016.
[31] S. M. Wong, “Residual Stress Measurements on Chromium Films
by X-Ray Diffraction Using the sin2
ψ
Method,” Thin Solid Films,
Vol.53, pp. 65-71, 1978.
[32] Q. Luo and A. H. Jones, “High-Precision Determination of Resid-
ual Stress of Polycrystalline Coatings Using Optimised XRD-sin2
ψ
Technique,” Surface and Coatings Technology, Vol.205, pp. 1403-
1408, 2010.
[33] S. Kikuchi, T. Imai, H. Kubozono, Y. Nakai, A. Ueno, and K.
Ameyama, “Evaluation of Near-Threshold Fatigue Crack Propaga-
tion in Ti-6Al-4V Alloy with Harmonic Structure Created by Me-
chanical Milling and Spark Plasma Sintering,” Frattura ed Integrit`a
Strutturale, Vol.34, pp. 334-340, 2015.
[34] S. Kikuchi, T. Imai, H. Kubozono, Y. Nakai, M. Ota, A. Ueno, and
K. Ameyama, “Effect of Harmonic Structure Design with Bimodal
Grain Size Distribution on Near-Threshold Fatigue Crack Propaga-
tion in Ti–6Al–4V Alloy, Int. J. of Fatigue, Vol.92, pp. 616-622,
2016.
[35] R. K. Nalla, B. L. Boyce, J. P. Campbell, J. O. Peters, and R. O.
Ritchie, “Influence of Microstructure on High-Cycle Fatigue of Ti-
6Al-4V: Bimodal vs. Lamellar Structures,” Metallurgical and Ma-
terials Transactions A, Vol.33, pp. 899-918, 2002.
[36] B. L. Boyce and R. O. Ritchie, “Effect of Load Ratio and Maximum
Stress Intensity on the Fatigue Threshold in Ti-6Al-4V,” Engineer-
ing Fracture Mechanics, Vol.68, pp. 129-147, 2001.
[37] H. Oguma and T. Nakamura, “The Effect of Stress Ratios on Very
High Cycle Fatigue Properties of Ti-6Al-4V,” Key Engineering Ma-
terials, Vol.261-263, pp. 1227-1232, 2004.
[38] JSMS Committee on Fatigue of Materials and JSMS Committee
on Reliability Engineering, Standard evaluation method of fatigue
reliability for metallic materials -Standard Regression Method of
SNCurves- [JSMS-SD-6-08], The Society of Materials Science
Japan, 2008.
[39] M. Haddad, R. Zitoune, F. Eyma, and B. Castani´e, “Influence of
Machining Process and Machining Induced Surface Roughness on
Mechanical Properties of Continuous Fiber Composites,” Experi-
mental Mechanics, Vol.55, No.3, pp. 519-528, 2015.
[40] M. Haddad, R. Zitoune, H. Bougherara, F. Eyma, and B. Cas-
tani´e, “Study of Trimming Damages of CFRP Structures in Func-
tion of the Machining Processes and Their Impact on the Mechan-
ical Behavior,” Composites Part B: Engineering, Vol.57, pp. 136-
143, 2014.
[41] T. Morita, “Effects of shot peening condition on surface properties
and fatigue strength of Titanium,” Mechanical Surface Tech, Vol.10,
pp. 36-38, 2016.
[42] K. N. Smith, T. H. Topper, and P. Watson, “A Stress-Strain Function
for the Fatigue of Metals (Stress-Strain Function for Metal Fatigue
Including Mean Stress Effect),” J. of Materials, Vol.5, pp. 767-778,
1970.
[43] S. Heinz, F. Balle, G. Wagner, and D. Eifler, “Analysis of Fatigue
Properties and Failure Mechanisms of Ti6Al4V in the Very High
Cycle Fatigue Regime Using Ultrasonic Technology and 3D Laser
Scanning Vibrometry,” Ultrasonics, Vol.53, pp. 1433-1440, 2013.
[44] D. F. Neal and P. A. Blenkinsop, “Internal Fatigue Origins in
α
-
β
Titanium Allols,” Acta Metallurgica, Vol.24, No.1, pp. 59-63,
1976.
Name:
Shoichi Kikuchi
Affiliation:
Department of Mechanical Engineering, Gradu-
ate School of Engineering, Kobe University
Address:
1-1 Rokkodai-cho, Nada-ku, Kobe 6578501, Japan
Brief Biographical History:
2008- GCOE Researcher, Keio University
2010- Assistant Professor, Ritsumeikan University
2013- Visiting Researcher, University of Kaiserslautern
2014- Assistant Professor, Kobe University
Main Works:
S. Kikuchi, T. Imai, H. Kubozono, Y. Nakai, M. Ota, A. Ueno, and K.
Ameyama, “Effect of harmonic structure design with bimodal grain size
distribution on near-threshold fatigue crack propagation in Ti-6Al-4V
Alloy, Int. J. of Fatigue, Vol.92, pp. 616-622, March, 2016.
Membership in Academic Societies:
Society of Materials Science, Japan (JSMS)
Japan Society of Mechanical Engineers (JSME)
Japan Institute of Metals and Materials (JIM)
Japan Society for Abrasive Technology (JSAT)
Name:
Yuki Nakamura
Affiliation:
Department of Mechanical Engineering, Na-
tional Institute of Technology, Toyota College
Address:
2-1 Eisei-cho, Toyota 471-8525, Japan
Brief Biographical History:
2011- Assistant Professor, National Institute of Technology, Toyota
College
2014- Lecturer, National Institute of Technology, Toyota College
Main Works:
Y. Nakamura, T. Sakai, H. Hirano, and K. S. Ravi Chandran, “Effect of
Alumite Surface Treatments on Long-life Fatigue Behavior of a Cast
Aluminum Alloy in Rotating Bending,” Int. J. of Fatigue, Vol.32,
pp. 621-626, March, 2010.
Membership in Academic Societies:
Society of Materials Science, Japan (JSMS)
Japan Society of Mechanical Engineers (JSME)
Japan Society for Abrasive Technology (JSAT)
Japan Society for Design Engineering (JSDE)
Material Testing Research Association of Japan (MTRAJ)
Int. J. of Automation TechnologyVol.11 No.6, 2017 923
Kikuchi, S. et al.
Name:
Koichiro Nambu
Affiliation:
Toyota Technological Institute
Address:
2-12-1 Hisakata, Tempaku-ku, Nagoya 468-8511, Japan
Brief Biographical History:
2011- Assistant Professor, National Institute of Technology, Suzuka
College
2017- Assistant Professor, Toyota Technological Institute
Main Works:
K. Nambu, K. Monda, K. Inagaki, Y. Maeyama, and S. Kikuchi, “Effect
of Hardness Ratio on the Behavior of Plastic Deformation in Various
Metallic Materials Treated with Fine Particle Peening,” The 30th Int. Conf.
on Surface Modification Technology (SMT30) proc., 2016.
Membership in Academic Societies:
Society of Materials Science, Japan (JSMS)
Japan Society of Mechanical Engineers (JSME)
Japan Society for Abrasive Technology (JSAT)
Japan Society for Heat Treatment (JSHT)
Japanese Society of Tribologists (JAST)
Name:
Toshikazu Akahori
Affiliation:
Department of Materials Science and Engineer-
ing, Faculty of Science and Technology, Meijo
University
Address:
1-501 Shiogamaguchi, Tempaku-ku, Nagoya 468-8502, Japan
Brief Biographical History:
2000- Assistant Professor, Toyohashi University of Technology
2006- Assistant Professor, Tohoku University
2008- Associate Professor, Tohoku University
2010- Associate Professor, Miejo University
Main Works:
Y. Oguchi, T. Akahori, T. Hattori, H. Fukui, and M. Niinomi, “Change in
Mechanical Strength and Bone Contactability of Biomedical Titanium
Alloy with Low Young’s Modulus Subjected to Fine Particle Bombarding
Process,” Materials Transactions, Vol.56, pp. 218-233, November, 2015.
Membership in Academic Societies:
Japan Institute of Metals and Materials (JIM)
Society of Materials Science, Japan (JSMS)
Japan Society of Mechanical Engineers (JSME)
924 Int. J. of Automation TechnologyVol.11 No.6, 2017
... In this section, we compare the fatigue properties of CP titanium having a heterogeneous nitrogen diffusion phase with those of a bulk titanium alloy. The results are plotted as S-N diagrams in Fig. 15 for un-nitrided bulk Ti6Al4V 42) and CP titanium compacts that were made from powder gasnitrided at 823 K and that were then consolidated at 1473 K, which showed the highest fatigue limit in this study. A fatigue limit is clearly observed for each specimen; therefore, the S-N curves can be drawn using a bilinear S-N model with a fatigue limit (JSMS standard regression models). ...
... un-nitrided bulk Ti6Al4V: 42) log 10· a ¼ À35:196 log 10ðN f Þ þ 561:53; ...
... Fig. 15 Results of four-point bending fatigue tests for sintered CP titanium (sintering temperature: 1473 K) made from powder gas-nitrided at 823 K and for un-nitrided bulk Ti6Al4V alloy. 42) ...
Article
The purpose of this study is to develop commercially pure (CP) titanium having a higher fatigue strength than titanium alloys developed via heterogeneous nitrogen diffusion. The microstructure of CP titanium having a heterogeneous nitrogen diffusion phase, which was fabricated by consolidating gas-nitrided powders, was characterized, and its fatigue properties were examined. The nitrogen content and hardness of CP titanium compacts having a heterogeneous nitrogen diffusion phase increased with increasing powder gas nitriding temperature and sintering temperature. The fatigue limit and fatigue life of CP titanium compacts increased with increasing sintering temperature and with decreasing powder gas-nitriding temperature. In particular, CP titanium having a heterogeneous nitrogen diffusion phase that is fabricated by high-temperature sintering of powders treated with low-temperature nitriding has a higher fatigue limit than un-nitrided bulk Ti–6Al–4V alloy. The fatigue limit of CP titanium can be controlled by optimizing the powder gas nitriding and sintering temperatures. Fig. 12 (a) Optical micrograph and (b) nitrogen map of sintered compact fabricated from powder gas-nitrided at 873 K (sintering temperature: 1273 K) tested at σa = 240 MPa and N = 7 × 10³ cycles. Fullsize Image
... The alumina blasting treatment removed the metal particles, reducing Ra and Rz-that is, it was expected that the strength would increase as the stress concentration was reduced by smoothing the surface [21]. It was also thought that there could be a peeling effect due to the alumina blast treatment of the surface [19,28]. However, it was confirmed that oxygen and aluminum adhered to the sandblast-treated surface. ...
Article
Full-text available
Mandibular reconstruction using a titanium mesh tray and autologous bone is a common procedure in oral and maxillofacial surgery. However, there can be material problems—such as broken titanium mesh trays—which may undermine long-term functionality. This study was designed to investigate the optimal conditions for a titanium mesh tray with an ideal mandibular shape and sufficient strength, using computer-aided design, computer-aided manufacturing technology, and electron beam additive manufacturing. Specimens were prepared using Ti-6Al-4V extra low interstitial titanium alloy powder and an electron beam melting (EBM) system. The mechanical strength of the plate-shaped specimens was examined for differences in the stretch direction with respect to the stacking direction and the presence or absence of surface treatment. While evaluating the mechanical strength of the tray-shaped specimens, the topology was optimized and specimens with a honeycomb structure were also verified. Excellent mechanical strength was observed under the condition that the specimen was stretched vertically in the stacking direction and the surface was treated. The results of the tray-shaped specimens indicated that the thickness was 1.2 mm, the weight reduction rate was 20%, and the addition of a honeycomb structure could withstand an assumed bite force of 2000 N. This study suggests that the EBM system could be a useful technique for preparing custom-made titanium mesh trays of sufficient strength for mandibular reconstruction by arranging various manufacturing conditions.
... In recent years, it has been reported that FPP treatment can form a surface-modified layer enriched with the colliding particles of carbon steel [14], titanium [15], chromium [16], and carbon-black/steel [17] at room temperature. In our previous studies [18][19][20], a HAp transfer layer was formed by FPP using HAp particles on CP titanium, Ti-6Al-4V alloy, and Ti-29Nb-13Ta-4.6Zr alloy without any change in the crystalline structure of HAp. ...
Article
Full-text available
Fine particle peening (FPP) using hydroxyapatite (HAp) shot particles was performed to improve the fatigue strength and form a HAp transfer layer on a beta titanium alloy (Ti–22V–4Al). The surface microstructures of the FPP-treated specimen were characterized using scanning electron microscopy, micro-Vickers hardness testing, energy dispersive X-ray spectrometry, X-ray diffraction, and electron backscattered diffraction. A HAp transfer layer with a thickness of 5.5 μm was formed on the surface of the Ti–22V–4Al specimen by FPP. In addition, the surface hardness of the Ti–22V–4Al was increased, and high compressive residual stress was generated on the specimen surface by FPP. Rotating bending fatigue tests were performed at room temperature in laboratory air over a wide cycle-life region (103–109 cycles). In the long cycle-life regime, the fatigue strength at 107 cycles of the FPP-treated specimen became higher than that of the untreated specimen. This result is attributed to the formation of a work-hardened layer with high compressive residual stress by FPP. However, the fatigue strength was not improved by FPP in the short cycle-life regime, because fatigue cracks were initiated at surface defects formed during the FPP process. The fatigue fracture mode of the FPP-treated specimens shifted from surface-initiated fracture to subsurface-initiated fracture at a stress amplitude level of 600 MPa.
Article
Conventional nitriding improves the tribological properties of titanium alloys; however, reduces their fatigue strength owing to grain-coarsening. The purpose of this study is to examine the effect of the nitrided-fine particle peening (N-FPP) on the formation of nitrided layer and fatigue properties of Ti-6Al-4V alloy. This approach forms the nitrided layer on Ti-6Al-4V alloy by bombardment with nitrided commercially pure (CP) titanium fine particles. Plasma nitriding and gas nitriding were performed at 873 or 973 K to form a nitrided layer on the surface of CP-titanium fine particles, and then N-FPP was performed for 1, 10, 30 s in air at room temperature. Nitrided layer with high hardness and compressive residual stress could be found on the N-FPP treated surface of Ti-6Al-4V alloy. Nitrogen concentration in nitrided layer tended to increase with the particle nitriding temperature and N-FPP treatment time. Fatigue life and fatigue limit of the N-FPP treated specimen under four-point bending fatigue tests were higher than those of conventional nitrided one, whereas were lower than those of the un-peened specimen. This was because a fatigue crack in the N-FPP treated specimen was initiated from the surface dent formed by N-FPP due to the stress concentration. In contrast, FPP using un-nitrided CP-titanium fine particles increased the fatigue limit of Ti-6Al-4V alloy.
Article
Single-phase equiatomic high-entropy alloy (HEA); CrMnFeCoNi, exhibits high strength and ductility at room temperature and low temperature due to the effect of twinning. In this study, a concept of a bimodal microstructure design for HEA using powder metallurgy was proposed. Microstructures of the sintered compact fabricated from HEA powder mechanically-milled using SUJ2 balls were analyzed by EBSD. A network structure of fine grains (Shell), which surrounded the coarse-grained structure (Core), was formed in the HEA compact. The hardness of the HEA with bimodal microstructure were higher than the compact fabricated from as-received HEA powder, and the network structure showed high hardness than Core phase. Furthermore, W-rich surface layer was formed on the HEA powder mechanically-milled using WC-Co balls owing to the transfer of WC-Co balls to the surface of HEA powder during mechanical milling. A network structure of the W-rich phase, which surrounded the HEA phase, was formed in the compact fabricated from the mechanically-milled HEA powder. In particular, the hardness of compact fabricated from HEA powder mechanically-milled using WC-Co balls was high due to the high concentration of tungsten and formation of Shell phase. These results indicate that developed method can control the microstructure of HEA; grain size and elementary diffusion distributions.
Article
Full-text available
Multifunctional surfaces are required to design safe engineering products for human lives. Heating in a nitrogen atmosphere (nitriding) improves the tribological properties but reduces the strength of titanium (Ti) alloys owing to grain coarsening. A rapid nitriding method for Ti alloys forms the nitrided layer on the surface of a Ti alloy by bombarding with commercially pure Ti fine particles with a nitrided phase at room temperature within a short period. Furthermore, fine grains of Ti alloy are formed in the nitrided layer because of the impact of the Ti particles. These results reveal that this room‐temperature method resolves the trade‐off between the rapid formation of a nitrided layer and the suppression of grain coarsening for Ti alloys. The mechanical and tribological properties are important for titanium alloys, which are used as a biomaterial. Inspiration is taken from the transfer of fine particles to achieve rapid nitriding without heating. The developed room‐temperature methods result in the formation of a nitrided layer within a short period. The proposed method has the potential to improve various properties in other metallic materials.
Article
A beta-titanium alloy, which has good biocompatibility and low Young’s modulus, is expected to use for biomedical applications. In this study, in order to investigate the near-threshold fatigue crack propagation in a beta-titanium alloy (Ti-29Nb-13Ta-4.6Zr; TNTZ) with low Young’s modulus, stress intensity factor decreasing tests were conducted under the force ratios from 0.1 to 0.8 in air at room temperature. After testing, crack profiles were observed by scanning electron microscopy, and microstructures around crack profiles were analyzed using electron backscatter diffraction to discuss the mechanism of fatigue crack propagation. The crack growth rate in the solution-treated TNTZ followed by aging were constantly higher at comparable stress intensity range levels, and its threshold stress intensity ranges were lower compared to the only solution-treated TNTZ. This is attributed to the reduction of the opening stress intensity factor resulting from the formation of the alpha-phase by aging. However, the effect of microstructure on fatigue thresholds in TNTZ was disappeared by eliminating the crack closure phenomenon.
Article
During aluminum die-casting, tensile residual stress accumulates on the cavity surface of the die by repeated heating and cooling processes. Recently, to improve productivity, dies with high cycle and longer life have become necessary, and reduction or removal of tensile residual stress can be used to prevent heat cracks that cause mold fracture. Heat treatment is often used for residual stress reduction but a more efficient residual stress reduction method that can be carried out with simpler equipment is required. In this study, the relationship between the residual stress after forced vibration and the amplitude at the time of excitation is investigated by mechanical vibration of the SKD61 die materials and the die-casting mold through the application of forced vibration by an eccentric motor. Residual stress on the surface of each test plate treated by the heat treatment and the surface of mold cavity after excitation is evaluated by the X-ray residual stress measurement. It was found that the residual strain after excitation accumulated in compression as the amplitude of oscillation of the specimen became negative. Residual stress in the excitation direction of the specimens increased in the compression direction due to the excitation, demonstrating the effective stress reduction by the excitation method.
Article
Full-text available
The effect of a fine particle peening (FPP) on atmospheric oxidation behavior and tribological properties of Ti-6Al-4V alloy was evaluated. Surface microstructures of oxidized specimens pre-treated with FPP were characterized using scanning electron microscope (SEM), energy dispersive spectrometry (EDS), glow discharge optical emission spectrometry (GDOES) and X-ray diffraction (XRD). The oxide layer formed on the oxidized specimen pre-treated with FPP was thicker than that on the oxidized-only specimen, because the microstructure induced by FPP facilitated the diffusion of oxygen and aluminum elements during the oxidation process. As results of reciprocating sliding wear tests, width of wear track on the oxidized specimen pre-treated with FPP was shallower compared to the oxidized-only specimen. Moreover, the oxide layer formed at the oxidized-only surface was delaminated during tests, otherwise there was no delamination at the oxidized surface pre-treated with FPP. This was because the surface oxide layer exhibited good interface adherence due to the existence of a thick oxygen solid solution layer. These results indicate that the modified layer created by the combination process of FPP and atmospheric oxidation is effective to improve the wear resistance of titanium alloys.
Article
Full-text available
Titanium alloy (Ti-6Al-4V) having a bimodal “harmonic structure”, which consists of coarsegrained structure surrounded by a network structure of fine grains, was fabricated by mechanical milling (MM) and spark plasma sintering (SPS) to achieve high strength and good plasticity. The aim of this study is to investigate the near-threshold fatigue crack propagation in Ti-6Al-4V alloy with harmonic structure. Ti-6Al-4V alloy powders were mechanically milled in a planetary ball mill to create fine grains at powder’s surface and the MM-processed powders were consolidated by SPS. K-decreasing fatigue crack propagation tests were conducted using the DC(T) specimen (ASTM standard) with harmonic structure under the stress ratios, R, from 0.1 to 0.8 in ambient laboratory atmosphere. After testing, fracture surfaces were observed using scanning electron microscope (SEM), and crack profiles were analyzed using electron backscatter diffraction (EBSD) to discuss the mechanism of fatigue crack propagation. Threshold stress intensity range, ΔKth, of the material with harmonic structure decreased with stress ratio, R, whereas the effective stress intensity range, ΔKeff, showed constant value for R lower than 0.5. This result indicates that the influence of the stress ratio, R, on ΔKth of Ti- 6Al-4V with harmonic structure can be concluded to be that on crack closure. Compared to the compact prepared from as-received powders with coarse acicular microstructure, ΔKth value of the material with harmonic structure was low. This was because the closure stress intensity, Kcl, in the material with harmonic structure was lower than that of the coarse-grained material due to the existence of fine grains. In addition, the effects of the grain size on the fatigue crack propagation behaviors of Ti-6Al-4V alloy were investigated for the bulk homogeneous material. The effects of the stress ratio and the grain size on the fatigue crack propagation of the material with harmonic structure were quantified.
Article
The emergence of biodegradable bone substitute materials promises new breakthroughs in the field of tissue engineering. However, such materials like Hydroxyapatite and Calcium Polyphosphate (CPP) are relatively new in the world of manufacturing. Reliable processing data is required in order optimize the production of biomedical implants made from these materials. This paper investigates the machining characteristics of CPP.Milling studies are conducted to determine the suitable cutting speeds and conditions which produce accurately machined porous surfaces. This is required for the CPP material to fulfill its biomedical function. A force model is also developed by identifying the cutting coefficients, which helps to predict and limit the machining forces, thereby avoiding unwanted chipping of the implant. Results of these studies have been successfully incorporated into planning the 5-axis machining of a CPP construct for a tissue engineered tibial-plateu (lower knee joint) implant.
Article
A recently suggested stress- strain function governing fatigue is extended to include the effect of mean stress. This function is probably valid, at least in the engineering sense, for the crack initiation stage and the early part of crack propagation. A procedure for cumulative damage summation in terms of this function is presented. The proposal is tested against various metals (steels and aluminum alloys) and loading conditions.
Article
First, in this paper, a new atmospheric-controlled induction heating and fine particle peening treatment system (vacuum AIH-FPP system) which reduces the oxygen concentration in the chamber to the order of ppm, much less than a conventional processing apparatus was presented. Next, in order to examine the effect on the formation of the surface modified layer of (i) mixing hard particles, (ii) the processing temperature, and (iii) the particle velocity, carbon steel AISI 1045 was treated with this system in conjunction with high-frequency induction heating, by peening Cr particles and mixed particles of Cr and high-speed tool steel. From the observation results by a scanning electron microscope and an energy dispersive X-ray spectrometer, it is clear that for the formation of a Cr diffused layer, using a mixture of Cr particles and high-speed tool steel particles is important. The treatment must be conducted at a higher temperature of approximately 1273 K to form a Cr diffused layer. Furthermore, by increasing the particle velocity, a thicker Cr transfer layer is formed at the surface under process. Therefore, an increased particle velocity accelerates the transfer of Cr. This Paper was Originally Published In Japanese in J. Japan Inst. Met. Mater. 79 (2015) 491–496. In order to establish an effective method of the surface treatment proposed in this paper, some parts of the contents were revised and Figure 10 was added. The sentences of abstract, conclusions, and references were slightly modified. Two contributed authors were also added.
Article
Titanium alloy (Ti–6Al–4V) with a bimodal harmonic structure, which consists of a coarse-grained structure surrounded by a network structure of fine equiaxed grains, has been fabricated by sintering mechanically-milled powders to achieve high strength and good plasticity. To investigate the near-threshold fatigue crack propagation in the harmonic structured Ti–6Al–4V alloy, K-decreasing tests are conducted on disk-shaped compact specimens (ASTM standard) under stress ratios R from 0.1 to 0.8 with a constant-R loading regime in a laboratory atmosphere. The fracture surfaces are observed using scanning electron microscopy (SEM), and crack profiles are analysed using electron backscatter diffraction (EBSD) to discuss the mechanism of the fatigue crack propagation. The crack growth rates da/dN in the harmonic structured material are constantly higher than those in a material with coarse acicular microstructure under comparable stress intensity range ΔK, while the fatigue thresholds ΔKth are lower. This is attributed to a decrease in the magnitude of roughness-induced crack closure and the effective stress intensity range ΔKeff,th in the harmonic structured Ti–6Al–4V alloy due to the presence of fine grains. Furthermore, in some areas, fatigue cracks do not propagate in the coarse-grained structure with higher fatigue crack growth resistance, but they preferentially propagate across the network structure of fine grains in the harmonic structure.
Article
Fine particle peening (FPP) using hydroxyapatite (HAp) shot particles was introduced to form the HAp surface layer and improve the fatigue properties of commercially pure (CP) titanium. The surface microstructure of the FPP-treated specimens was characterized using a micro-Vickers hardness tester, optical microscopy, scanning electron microscopy (SEM), energy dispersive X-ray spectrometry (EDX), X-ray diffraction (XRD), and non-contact scanning white light interferometry. FPP could create a HAp layer on the surface of CP titanium within a relatively short time (1 s) by shot particle transfer. In addition, FPP increased the surface hardness and generated compressive residual stress at the treated surface. Four-point bending fatigue tests were performed at a stress ratio of 0.1 in air at room temperature to examine the effect of FPP using HAp shot particles on the fatigue properties of CP titanium. It was found that the fatigue limit for the FPP-treated specimen was higher than that for the unpeened specimen. The fatigue fracture mechanism for the CP titanium treated with FPP was discussed from the viewpoint of fractography. The HAp layer remained on the surface without delamination after the fatigue tests.
Article
In this study, we developed a surface modification technology for implants using commercially pure (cp) Ti. The technology used in this study leads to reduction in the time required for adhesion between bone and surfaces of implants. The existence ofmicroasperities and oxide layers is important to induce calcium phosphate precipitation and bone formation activity of osteoblasts. In addition, we focused on nanosecondpulsed laser treatment as a method to create both microasperities and oxide layers. First, we observed surface morphologies formed by laser treatment. An oxide layer with high oxygen concentration and microasperities on the order of 10 nm to 10 μmwere produced. Moreover, the OH groups were created on the laser-treated surface. Second, by culturing osteoblasts on the laser-treated cp Ti surface, its effects on cell shape, proliferation, and activity of bone formation were evaluated. Even though cell proliferation was at a comparable level in these two surfaces, the ALP activity per cell number was improved by about four times in the laser-treated surface compared with that in the polished surface. On the laser-treated cp Ti surface, it was considered that the bone formation activity of osteoblasts was promoted without inhibiting cell proliferation. From the results of this study, it is possible to conclude that by treating cp Ti surfaces with a laser, a surface with good cytocompatibility can be created.