Sie sind auf Seite 1von 7

Materials Science and Engineering C 50 (2015) 5258

Contents lists available at ScienceDirect

Materials Science and Engineering C


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

First principles theoretical investigations of low Young's modulus beta


TiNb and TiNbZr alloys compositions for biomedical applications
Rajamallu Karre a, Manish K. Niranjan b, Suhash R. Dey a,
a
b

Dept. of Materials Science and Metallurgical Engineering, Indian Institute of Technology Hyderabad, Yeddumailaram 502205, Telangana, India
Dept. of Physics, Indian Institute of Technology Hyderabad, Yeddumailaram 502205, Telangana, India

a r t i c l e

i n f o

Article history:
Received 17 September 2014
Received in revised form 12 December 2014
Accepted 16 January 2015
Available online 24 January 2015
Keywords:
Alloy design
Biomedical
Density functional theory
Ab-initio calculations
Elastic properties
Phase stability-prediction

a b s t r a c t
High alloyed -phase stabilized titanium alloys are known to provide comparable Young's modulus as that to the
human bones (~30 GPa) but is marred by its high density. In the present study the low titanium alloyed compositions of binary TiNb and ternary TiNbZr alloy systems, having stable -phase with low Young's modulus are
identied using rst principles density functional framework. The theoretical results suggest that the addition of
Nb in Ti and Zr in TiNb increases the stability of the -phase. The -phase in binary TiNb alloys is found to be
fully stabilized from 22 at.% of Nb onwards. The calculated Young's moduli of binary -TiNb alloy system are
found to be lower than that of pure titanium (116 GPa). For Ti25(at.%)Nb composition the calculated Young's
modulus comes out to be ~80 GPa. In ternary TiNbZr alloy system, the Young's modulus of Ti25(at.%)Nb
6.25(at.%)Zr composition is calculated to be ~50 GPa. Furthermore, the directional Young's moduli of these two
selected binary (Ti25(at.%)Nb) and ternary alloy (Ti25(at.%)Nb6.25(at.%)Zr) compositions are found to be
nearly isotropic in all crystallographic directions.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Titanium and its alloys have attracted considerable attention and
interest in recent years due to their potential biomedical applications
[1,2]. These alloys possess high specic strength, superior biocompatibility, excellent wear and corrosion resistance in biological environment
and provide adequate mechanical properties when compared to other
metallic biomaterials [3,4]. Pure Ti shows allotropic phase transformation at 1154 K, changing from low temperature hcp phase to high
temperature bcc phase [5]. This -transus temperature in titanium
alloys is dependent on the alloying/interstitial elements. Though titanium metal and alloys have comparatively smaller elastic moduli than
the other metallic biomaterials, it is still quiet higher than that of the
human bone. This is important as the difference in Young's moduli of
bone and metallic biomaterial implant leads to stress shielding effect
which results into bone resorption over a long period of time [6,7].
However, this effect can be minimized in titanium alloys by introducing
-phase microstructure which has lower elastic modulus than the
-titanium alloys [8]. The stabilizing elements can be classied mainly
into two categories: isomorphous (e.g. Mo, Nb, Ta, and Zr) and eutectoid
(e.g. Cr, Mn, Fe, and Co) [9]. In particular, the -isomorphous titanium
alloys with nontoxic and allergy-free alloying elements have been

Corresponding author.
E-mail address: suhash@iith.ac.in (S.R. Dey).

http://dx.doi.org/10.1016/j.msec.2015.01.061
0928-4931/ 2015 Elsevier B.V. All rights reserved.

found very promising due to their low Young's modulus equivalent to


that of the bone which in turn enhances the mechanical biocompatibility [1012]. The volume fraction of phase can be enhanced in titanium
alloys by the introduction of stabilizing transition-metal (TM) elements such as Mo, V, Nb, Cr, and Ta [7]. However, an increasing content
of stabilizer alloying elements in titanium may give rise to other
phases such as hexagonal , orthorhombic and hexagonal omega
() depending upon heat treatment. Thus, the content of TM alloying
elements in titanium should be kept as low as possible and must be sufcient enough to retain the phase. In addition, these stabilizers
(alloying elements) should be non-toxic and non-allergic to be qualied
for biomaterials [13,14]. Through these criteria niobium (Nb), zirconium (Zr) and tantalum (Ta) are favorable alloying elements for titanium
over others [12].
In recent years, computational material design based on rst principles quantum mechanical methods have emerged as an important cost
effective tool to design material properties [15,16]. Although biocompatible Ti alloys have been investigated experimentally, very few theoretical studies of them have been reported. Recently Raabe et al. [17]
have reported directional Young's moduli for binary Ti alloys (TiNb
and TiMo) up to 50 at.% with a composition variation of 6.25 at.%,
whereas Ikehata et al. [18] have reported polycrystalline Young's moduli of binary Ti1 xMx (M = V, Nb, Mo, and W) alloys for x = 0.0, 0.25,
0.5, 0.75 and 1 with specic structural conguration.
However, in calculations performed by Ikehata et al. the compositional variation within these binary alloy systems is large (variation 25

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

at.%). Further, not all possible structural congurations were included in


their calculations.
In this article, we use rst-principles calculations within density
functional framework to study binary TiNb and ternary TiNbZr
systems with a minimum composition variation of 6.25 at.% for all
possible structural congurations. In particular we study the stability
of phase in binary TiNb and ternary TiNbZr alloys with respect to
phase and subsequently determine lowest polycrystalline Young
modulus compositions within these two alloy systems.
2. Calculation methodology
The rst principles calculations are performed within density
functional framework [19] with the projected augmented wave
(PAW) [20] method as implemented in the VASP code [21]. The
PerdewBurkeErnzerhoff (PBE) [22] form of the generalized gradient approximation is used to approximate exchange and correlation
energy. The electron wave functions are expanded in a standard
plane wave basis set with a kinetic energy cutoff of 400 eV. The
calculations are performed using the (18/n1) (18/n2) (18/n3)
and (20/n1) (20/n2) (18/n3) MonkhorstPack k-point meshes
for the n1 n2 n3 cubic bcc and hcp supercell respectively. The
n1, n2, and n3 are the numbers of unit cells taken in x, y, and z directions respectively. The calculations are converged to 10 7 eV/cell
and the structures are relaxed until the largest force becomes
less than 10 3 eV/. In the present study we focus on binary
TiNb and ternary TiNbZr systems. Zirconium is added as ternary element in TiNbZr because of its superior biocompatibility
and its composition is kept xed at 6.25 at.%. The calculations are
performed using a sixteen atom supercell for binary TiNb as well
as ternary TiNbZr system. The bcc and hcp unit cells of pure
metals are assumed to have Im3m and P6 3 /mmc space groups
respectively.

where E el is the energy due to electronic excitations and S el is the


bare electronic entropy and are given by
EF

Eel V; T f   n; V d  n; V d
f f ln f 1f ln 1f g  n; V d

In the equilibrium state of a crystal or an alloy system the Helmholtz


free energy

SConfig x kB  x ln x 1x  ln 1x:

In our calculations we have not included the vibrational contribution to the free energy due to limitations of computational resources. It may be noted that vibrational contribution to the free
energy can be sizeable only at elevated temperatures (around
1273 K). At body temperature (310 K) the primary contribution to
temperature dependent free energy term comes from congurational entropy [23]. Furthermore, the contribution of vibrational entropy
to the formation energy in alloy system is expected to be much
smaller than that of congurational entropy due to mutual cancelation of vibrational entropy contributions. We have also not included contribution of electronic excitations to the free energy since
temperature scale relevant to this study is well below the electronic
energy scale rendering this contribution very small. It should be
noted that the calculation of electronic excitation contribution to
the free energy is straightforward. Thus in the present work we
studied the thermodynamic stability of an alloy system using following expression
T  Sconfig

bcc/hcp
where E form
is the formation energy of TiNb alloy system and
given by

bcc=hcp

F V; T UTS

where f and n(, V) are Fermi-distribution function and density of


electronic states respectively.
The congurational entropy of a binary alloy system AxB1 x, in ideal
mixing approximation is given by

bcc=hcp

3.1. Free energy and thermodynamic stability

Z
Sel V; T kB

F x; T Eform

3. Theoretical details

53

Eform

Ntotal

bcc=hcp

Tihcp

Nbbcc

Esupercell NTi  Ebulk N Nb Ebulk

at given temperature, T, and volume, V, is at minimum with respect


to variations of degrees of freedom such as electronic states and
atomic structure. At nite temperature, F contains various contributions such as vibrational, electronic and congurational and is given
by

where Ntotal, NTi and NNb are total number of atoms, number of Ti
bcc/hcp
atoms and number of Nb atoms in the supercell. Esupercell
is the energy of the bcc or hcp supercell. ETi(hcp)
and ENb(bcc)
are the energies
bulk
bulk
per atom of bulk Ti and Nb in hcp and bcc phases respectively.

F V; T EV F vib ; T F el V; T T  Sconfig

3.2. Elastic constants and Young's modulus

where E(V) is 0 K static contribution to the internal energy and is


easily calculated within ab-initio DFT framework. F vib and F el are
the vibrational and the electronic contributions respectively to the
free energy. Scong is the congurational entropy in the alloy system.
The Fvib in quasiharmonic approximation is given by
F vib V; T kB T


XX
!
q


19
8
0
!
<
=
hn q ; V
A
ln 2 sinh@
:
;
2kB T


!
q ; V is the frequency of nth phonon mode at the wave vec-

here n
!
tor q in the Brillouin zone. The Fel can be obtained from electronic
density of states and is given by
F el V; T Eel V; T T  Sel V; T

To calculate polycrystalline Young's modulus (E) we rst compute


elastic constants of the alloy systems. The Young's modulus (E) is
given by
E

9BG
3B G

10

where B and G are the polycrystalline bulk and shear moduli in the Hill's
approximation [24] and are given by
B

1
B BV ;
2 R

1
G GV
2 R

11

where subscripts R and V indicates bulk and shear moduli in Reuss and
Voigt approximations respectively and can be calculated from the elastic and the compliance constants. There are three, six and nine independent elastic constants for crystals in cubic, hexagonal and orthorhombic

54

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

symmetries. The bulk and shear moduli for an orthorhombic crystal are
given by
BV C 11 C 22 C 33 2C 12 2C 13 2C 23 =9
BR S11 S22 S33 2S12 2S13 2S23 
GV

1
3
C C 22 C 33 C 12 C 13 C 23 
C C 55 C 66 
15 11
15 44

12
13
14

GR 154S11 S22 S33 3S44 S55 S66 4S12 S13 S23 

15
where Cij and Sij are the components of elastic and compliance tensors.
The compliance constant matrix (S) and elastic constant matrix (C) are
related as S = C1. Further, the Young's moduli in each of the Cartesian
directions can be calculated from the elastic compliances and are given
by
1

Ex S11 ; Ey S22 ; Ez S33 :

16

The details of elastic constants calculations can be found in Ref. [25].


4. Results and discussions
4.1. Bcc () phase stability of binary TiNb and ternary TiNbZr alloy
systems
The formation energies of binary and ternary titanium alloy compositions are calculated using Eqs. (8) and (9). A comparative graph of
formation energy (meV/atom) of phase for binary and ternary
alloy system is plotted as a function of niobium content at different temperatures, 0 K, 310 K (37 C), and 1154 K (881 C) and are shown in
Figs. 1 & 2. As evident from Fig. 1(a), (b) and (c), the formation energy
of bcc () phase and hexagonal () phase in TiNb alloys decreases
and increases respectively as Nb content is increased. A decrease in formation energy indicates increase in stability of the phase. At T = 0 K, the
formation energy of bcc () phase crosses and becomes lower than that
of hexagonal () phase as the niobium content is increased by more
than 22 at.% (see Fig. 1(a)). This indicates that the bcc () phase
becomes more stable than hexagonal () phase beyond cross over
point. At higher temperatures, the formation energy of bcc () phase
of TiNb system further decreases due to entropy contribution
(Fig. 1(b) and (c)). However the cross-over point, i.e., minimum Nb content in TiNb system needed to stabilize the phase, varies only marginally and remains at ~ 22 at.%. Fig. 1(c) also shows the meniscus
which develops in formation energy curve for niobium content between
25 at.% and 100 at.% at large temperature. This is expected as the congurational entropy is highest for 50 at.% Nb content and vanishes when
Nb content is 0 at.% and 100 at.%. Furthermore, the product of T and
Scong is signicant for large T (see Eq. (8)).
The computed value of the minimum niobium content also called
critical concentration (~22 at.%) is in good agreement with experimentally obtained values of ~ 25 at.% [7]. The small difference between
experimental and theoretical results may be attributed to excluded
vibrational contribution to theoretical formation energy and mixing of
unwanted interstitial elements (oxygen mainly) during production of
experimental samples.
The stability of bcc phase in TiM alloys (M = transition element)
can also be understood from the ratio of valence electrons and number
of atoms (e/a) as reported in Ref. [26]. It has been shown that phase in
TiM alloys is stabilized if the e/a ratio is 4.2 or more [27]. In TiNb alloy
system, the e/a 4.2 starts from Ti22Nb onwards. Thus minimum niobium content of ~ 2225 at.% is expected to stabilize TiNb phase.
This is consistent with present DFT and reported experimental results
[28-30]. The phase stability of TiNb system with 25 at.% Nb content

Fig. 1. Free formation energy as a function of Nb concentration in bcc and hcp Ti


(a) T = 0 K; (b) T = 310 K; (c) T = 1154 K.

can also be understood using bond order (Bo) and d orbital energy levels
(Md) as discussed in Ref. [31]. The values of Bo and Md are tabulated for
all the TiNb and TiNbZr compositions in Table 1. Fig. 5 shows the
values of Bo and Md for all TiNb compositions superimposed on map
reported in Ref. [31] As can be seen the phase starts to become stable
for Nb content more than 25 at.%. The stability of the phase can also be
correlated with the shear modulus C = (C11 C12) / 2 which is a {110}
b110N shear, as reported in Ref. [32] and shown in Fig. 6. It is apparent
from the gure that the stability of phase increases with increase in
Nb content.
Saito et al. [33] have suggested that -Ti alloys can exhibit multifunctional properties provided electronic magic numbers such as e/a
(electron to atom ratio) = 4.24, Bo = 2.87, and Md = 2.45. The composition Ti25(at.%)Nb is nearly close to satisfying all these electronic numbers simultaneously and is expected to give such multifunctional

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

55

Table 1
The BoMd values of calculated binary and ternary Ti alloy compositions.
S. no.

Composition

Bo

Md

1
2
3
4
5
6
7

Binary
Ti6.25Nb
Ti12.5Nb
Ti18.75Nb
Ti25Nb
Ti31.25Nb
Ti37.5Nb
Ti50Nb

2.809
2.829
2.847
2.867
2.886
2.906
2.944

2.445
2.444
2.443
2.441
2.44
2.438
2.43

8
9
10
11

Ternary
Ti25Nb6.25Zr
Ti25Nb12.5Zr
Ti25Nb18.75Zr
Ti25Nb25Zr

2.88
2.904
2.922
2.94

2.471
2.502
2.532
2.563

Nb content for ternary system at various temperatures (0 K, 310 K (37


C), and 1154 K (881 C)). It is evident from Fig. 2(c) that the addition
of Zr (6.25 at.%. content) enhances the stability of the phase in ternary
system with niobium up to 70 at.% at T = 1154 K. Experimental reports
also suggest that stability of phase increases with the increase in Zr
content in TiNb alloys [38]. However, in our calculations the addition
of Zr reduces the stability of phase at T = 0 K (see Fig. 2(a)). This effect
may be attributed to the hcp crystal structure of the pristine zirconium.
4.2. Low Young's modulus of binary TiNb and ternary TiNbZr alloy
systems
To assess the accuracy of our calculations, we rst calculate the
Young's moduli of pure Ti, Nb and Zr in their bulk phases using different
k-meshes. The calculated and the experimental values are shown in
Fig. 3. As evident, the theoretical values are in good agreement with experimental values. The Young's moduli of TiNb alloy system in bcc
phase are calculated with an increase in niobium content by 6.25 at.%
until it reaches 50 at.%. For each composition, all possible element placement congurations are considered. In total sixty possible congurations are taken into account for bcc TiNb binary compositions. The
average Young's modulus and the Young's modulus of the most stable
conguration as function of Nb content are given in Fig. 4. As can be
seen the Young's modulus of TiNb alloy compositions gradually increases with increasing Nb content nally saturating at ~ 25 at.% Nb.
The Young's modulus of TiNb system with 6.25 at.% Nb content is
~40 GPa and is lowest among all compositions. Incidentally this Young's
modulus is close to the human bone's Young's modulus (cortical ~ 20
30 GPa) [39,40]. However as discussed earlier the TiNb composition
with 6.25 at.% Nb content is unstable (high formation energy) as
Fig. 2. -phase formation energy vs. Nb concentration in TiNb with 1/16 Zr (red
line) and without Zr (black line) at temperatures (a) T = 0 K, (b) T = 310 K, and
(c) T = 1154 K.

properties such as ultra-strength, super-elasticity, super-plasticity, and


ultra-low Young's modulus. The values of e/a and calculated density of
each alloy compositions are shown in Fig. 4.
Next we choose TiNb system with 25 at.% Nb content for further
ternary element addition. In ternary TiNbZr alloy system, the Zr is
added as ternary element because of its superior biocompatibility as
well as reduction ability of other unwanted phase formation such as
omega phase () [34,35]. Generally, omega phase () is known to increase the Young's modulus [36]. The addition of Zr also gives rise to
shape memory effects (SME) and superelastic behavior in TiNb alloy
system which is an additional advantage for biomedical applications
[37]. Fig. 2(a), (b) and (c) shows the formation energy as function of

Fig. 3. Young's modulus (in GPa) of pure elements (Ti, Nb and Zr).

56

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

Fig. 6. Variations of shear moduli (C, C44), bulk modulus (B) and Young's modulus
(Y) with the increase in Nb content.

Fig. 4. Young's modulus of bcc TiNb system as a function of Nb content. Red line indicates avg. Young's modulus whereas blue line indicates Young's modulus of lowest formation energy conguration.

phase (see Fig. 1). As shown in Fig. 4, the computed average Young's
modulus of Ti25(at.%)Nb system is 81 GPa whereas it is 69 GPa for conguration with lowest formation energy. The theoretical results are in
good agreement with those obtained experimentally for similar composition binary TiNb alloys [30,41,42].
It is already reported that low elastic modulus may results from low
bonding force. As explained in Ref. [43], the bonding force of titanium
alloys based on Bo, Md and Zeff parameters can be estimated as the
following:

bonding force

Zeff Bo
M 2d

17

Fig. 5. Extended BoMd diagram taken from Ref. [31] superimposed with calculated binary
Tix(at.%)Nb (x = 6.25, 12.5, 18.75, 25, 31.25, 37.5 and 50).

where Zeff , Bo and M d are the compositional average values of Zeff, Bo


and Md in a titanium alloy. Zeff is the effective ionic charge and can be
calculated difference between the atomic number of the element
(Z) and the Slater shielding constant (s). The Ti based alloy system in
bcc () phase shows lower bonding force than its hexagonal ()
phase structure and hence, lower Young's modulus is expected from
its beta phase.
It has been suggested that phase Ti alloys with low Young's
modulus for biomedical applications should exhibit low shear moduli
(C, C44), low bulk modulus (B) and low phase stability [44,45]. Our
calculations suggest that the composition Ti25(at.%)Nb indeed exhibits
low stabilization energy and lowest Young's modulus (Fig. 4) among
all the binary compositions. In order to understand the trend of Young's
modulus, we show shear moduli (C and C44), and bulk modulus (B) as
function of Nb content for all binary TiNb alloys in Fig. 6. It can be seen
that though Ti25(at.%)Nb composition with lowest formation energy
has high B and C values, it exhibits low Young's modulus due to low
C44 value. However it should be noted that the observed softening phenomenon of C44 at the starting of phase stability region has not been
understood yet.
To study the effect of Zr content in Ti25(at.%)Nb system on Young's
modulus, ternary Ti25NbxZr with x varying from 6.25, 12.5, 18.75 to
25 at.% with maximum possible structural congurations are studied.
The results are presented in Fig. 7. The addition of Zr in Ti25(at%)Nb
does not affect the e/a ratio (though Bo, and Md are changed; see
Table 1) but rather increases the density further. It has been suggested
that the higher Bo value should give rise to low Young's modulus
among least stable Ti alloys [41].

Fig. 7. Young's modulus of bcc Ti25(at.%)NbxZr system as a function of Zr content. Red


line indicates avg. Young's modulus whereas blue line indicates Young's modulus of lowest formation energy conguration. The e/a remains 4.25 for all the compositions.

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

57

As clear from Fig. 7, Ti25(at.%)Nb6.25(at.%)Zr has the lowest average Young's modulus (50 GPa average value) among all the ternary bcc
Ti25NbxZr systems. However Ti25(at.%)Nb6.25(at.%)Zr has
slightly increased energy as compared to Ti(at.%)25Nb alloys. These
calculations suggest that Ti25(at.%)Nb6.25(at.%)Zr composition
holds promise as the most suitable candidate in ternary TiNbZr
alloy system for real biomedical applications.
4.3. Elastic anisotropy in TiNb and TiNbZr alloy systems
Next we study Young's modulus of bcc phase TiNb and TiNbZr
compositions in different crystallographic directions. The Young's modulus in a direction with directions cosines l1, l2, and l3 is dened as [46]

4
2 2
2 2
4
2 2
4
2 2
E l1 S11 2l1 l2 S12 2l1 l3 S13 l2 S22 2l2 l3 S23 l3 S33 l2 l3 S44 18
1
2 2
2 2
l1 l3 S55 l1 l2 S66
:
Figs. 8 & 9 show directional elastic moduli of bcc () structure of
binary Tix(at.%)Nb (x = 6.25, 12.5, 18.75, 25, 31.25, 37.5 and 50) and
ternary Ti25(at.%)Nbx(at.%)Zr (x = 6.25, 12.5, 18.75 and 25) alloys
with lowest formation energies. The left panels of Figs. 8 and 9 show
the elastic modulus in [100], [110] and [111] directions. Among TiNb
compositions, Ti6.25(at.%)Nb alloy composition shows most anisotropic tendency. Contrary to the earlier reported anisotropic behavior
(A = C44 / C) of Ti25(at.%)Nb alloy [19], the present calculations
shows relatively isotropic behavior. This can be attributed to the increase in E100 (with increasing C value; see Fig. 6) and E110 values
with increase in Nb content in TiNb system which gradually comes
closer to E111 value. In ternary Ti25(at.%)Nbx(at.%)Zr alloy system,
the addition of Zr decreases the directional Young's modulus in b100N
direction. Overall the Young's modulus exhibits anisotropic behavior
in ternary Ti25(at.%)Nbx(at.%)Zr alloy system for higher Zr contents.
5. Conclusions
Stable bcc () phase Ti alloy compositions containing biocompatible
Nb, Zr alloying elements for biomedical applications are predicted using
rst-principles density functional theory. These compositions are found
to have comparable Young's modulus with that of human bone
(30 GPa). The stability of phase and Young's moduli (average as well
as directional) of TiNb and TiNbZr alloy systems are calculated

Fig. 9. Direction dependent Young's modulus of bcc Ti(25 at.%)Nbx(at.%)Zr system


with various Zr content. The plot at left side shows the Young's modulus along three
crystallographic directions (b100N, b110N and b111N) only.

using rst principles calculations. The TiNb alloy system are found to
have stable bcc () phase for Nb content 22 at.% and higher. Ti(25
at.%)Nb alloy composition is found to be the best binary composition
with ~80 GPa (average value) of Young's modulus. Further, ternary Zr
alloying is opted to mix with the Ti25 at.% Nb alloy composition, mainly to decrease the formation energy of -phase and to reduce its Young's
modulus. The computed average Young's modulus (~50 GPa) is found
to be lowest for Ti25(at.%)Nb6.25(at.%)Zr composition. Further,
the computed directional Young's moduli in 25 at.% Nb and Ti
25(at.%)Nb6.25(at.%)Zr alloy compositions are found to exhibit isotropic behavior.
Acknowledgments
The authors would like to thank High Performance Computing
(HPC) Center, IIT Hyderabad for computational calculations. The
authors are grateful to Department of Science and Technology, India
for nancial assistance availed through Indo-Japan joint DST-JSPS project (DST/INT/JSPS/P-177/2014).
References

Fig. 8. Direction dependent Young's modulus of bcc TiNb system with various Nb
content. The plot at left side shows the Young's modulus along three crystallographic
directions (b100N, b110N and b111N) only.

[1] M. Geethaa, A.K. Singh, R. Asokamani, A.K. Gogia, Ti based biomaterials, the ultimate
choice for orthopaedic implants a review, Prog. Mater. Sci. 54 (2009) 397425.
[2] H.J. Rack, J.I. Qazi, Titanium alloys for biomedical applications, Mater. Sci. Eng. C 26
(2006) 12691277.
[3] M. Long, H.J. Rack, Titanium alloys in total joint replacement a materials science
perspective, Biomaterials 19 (1998) 16211639.
[4] K. Wang, The use of titanium for medical applications in the USA, Mater. Sci. Eng. A
213 (1996) I34I37.
[5] C. Leyens, M. Peters (Eds.), Titanium and Titanium Alloys Fundamentals and Applications, Wiley-VCH, 2003.
[6] D.R. Sumner, T.M. Turner, R. Igloria, R.M. Urban, J.O. Galante, J. Biomech. 31 (1998)
909917.
[7] M. Niinomi, M. Nakai, Titanium-based biomaterials for preventing stress shielding
between implant devices and bone, Int. J. Biomater. (2011) http://dx.doi.org/
10.1155/2011/836587.
[8] D. Kuroda, M. Niinomi, M. Morinaga, Y. Kato, T. Yashiro, Design and mechanical
properties of new -type titanium alloys for implant materials, Mater. Sci. Eng. A
243 (1998) 244249.
[9] E.W. Collings, Physical Metallurgy of Titanium Alloys, ASM International, Materials
Park, OH, 1988.
[10] M. Niinomi, Mechanical biocompatibilities of titanium alloys for biomedical applications, J. Mech. Behav. Biomed. Mater. 1 (2008) 3042.
[11] M. Niinomi, Recent research and development in titanium alloys for biomedical
applications and healthcare goods, Sci. Technol. Adv. Mater. 4 (2003) 445454.
[12] A. Biesiekierski, J. Wanga, M.A. Gepreel, C. Wen, A new look at biomedical Ti-based
shape memory alloys, Acta Biomater. 8 (2012) 16611669.

58

R. Karre et al. / Materials Science and Engineering C 50 (2015) 5258

[13] H. Kawahara, S. Ochi, K. Tanetani, K. Kato, M. Isogai, Y. Mizuno, H. Yamamoto, A.


Yamagami, J. Jpn. Soc. Dent. Mater. Dev. 4 (1963) 65.
[14] S.G. Steinemann, in: G.D. Winter, J.L. Leray, K. de Goot (Eds.), Evaluation of Biomaterials, John Wiley & Sons, Ltd., 1980, p. 1.
[15] R. Liu, X. Zhou, F. Yang, Y. Yu, Combination study of DFT calculation and experiment
for photocatalytic properties of S-doped anatase TiO2, Appl. Surf. Sci. 319 (2014)
5059.
[16] W.A. Counts, M. Friak, D. Raabe, J. Neugebauer, Using ab-initio calculations in designing bcc MgLi alloys for ultra-lightweight applications, Acta Mater. 57 (2009)
6976.
[17] D. Raabe, B. Sander, M. Friak, D. Ma, J. Neugebauer, Theory-guided bottom-up design
of -titanium alloys as biomaterials based on rst principles calculations: theory
and experiments, Acta Mater. 55 (2007) 44754487.
[18] H. Ikehata, N. Nagasako, T. Furuta, A. Fukumoto, K. Miwa, T. Saito, First-principles
calculations for development of low elastic modulus Ti alloys, Phys. Rev. B 70
(2004) 174113.
[19] W. Kohn, L.J. Sham, Self-consistent equations including exchange and correlation
effects, Phys. Rev. 140 (1965) A1133.
[20] P.E. Blchl, Projector augmented wave method, Phys. Rev. B 50 (1994) 1795317979.
[21] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B 47
(1993) 1.
[22] J.P. Perdew, K. Burke, M. Ernzerhoff, Generalized gradient approximation made simple, Phys. Rev. Lett. 77 (1996) 3865.
[23] M.K. Niranjan, Ab-initio determination of thermodynamic properties of CoSi2,
Comput. Mater. Sci. 95 (2014) 517521.
http://dx.doi.org/10.1016/
j.commatsci.2014.08.017.
[24] G. Grimvall, Thermophysical Properties of Materials, North Holland, Amsterdam,
1999.
[25] M.K. Niranjan, First principles study of structural, electronic and elastic properties of
cubic and orthorhombic RhSi, Intermetallics 26 (2012) 150156.
[26] E.W. Collings, The Physical Metallurgy of Titanium Alloys, Metallurgia, Moscow,
1988. 223.
[27] A.V. Dobromyslov, V.A. Elkin, Martensitic transformation and metastable beta phase in
binary titanium alloys with d-metals of 46 periods, Scr. Mater. 44 (2001) 905910.
[28] C.M. Lee, C.P. Ju, J.H. Chern Lin, Structureproperty relationship of cast TiNb alloys,
J. Rehabil. 29 (2002) 314322.
[29] Y. Mantani, M. Tajima, Phase transformation of quenched martensite by aging in
TiNb alloys, Mater. Sci. Eng. A 438440 (2006) 315319.
[30] Y. Hon, J. Wang, Y. Pan, Composition/phase structure and properties of titaniumniobium, Alloys Mater. Trans. 44 (11) (2003) 23842390.
[31] M. Abdel-Hady, K. Hinoshita, M. Morinaga, General approach to phase stability and
elastic properties of -type Ti-alloys using electronic parameters, Scr. Mater. 55
(2006) 477480.

[32] C. Zener, Contributions to the theory of beta-phase alloys, Phys. Rev. 71 (1947)
846851.
[33] T. Saito, T. Furuta, J. Hwang, S. Kuramoto, K. Nishino, N. Suzuki, et al., Multifunctional
alloys obtained via a dislocation-free plastic deformation mechanism, Science 300
(2003) 464467.
[34] Y.L. Hao, S.J. Li, S.Y. Sun, R. Yang, Effect of Zr and Sn on Young's modulus and
superelasticity of TiNb-based alloys, Mater. Sci. Eng. A 441 (2006) 112118.
[35] X. Tang, T. Ahmed, H.J. Rack, Phase transformations in TiNbTa and TiNbTaZr
alloys, J. Mater. Sci. 35 (2000) 18051811.
[36] H. Matsumoto, S. Watanebe, N. Masahashi, S. Hanada, Composition dependence of
Young's modulus in TiV, TiNb, and TiVSn alloys, Metal. Mater. Trans. A 37A
(2006) 32393249.
[37] J.I. Kim, H.Y. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Shape memory characteristics of Ti22Nb(28)Zr(at.%) biomedical alloys, Mater. Sci. Eng. A 403 (2005)
334339.
[38] M. Abdel-Hady, H. Fuwa, K. Hinoshita, H. Kimura, Y. Shinzato, M. Morinaga, Phase
stability change with Zr content in -type TiNb alloys, Scr. Mater. 57 (2007)
10001003.
[39] M. Pithioux, P. Lasaygues, P. Chabrand, An alternative ultrasonic method for measuring the elastic properties of cortical bone, J. Biomech. 35 (2002) 961968.
[40] H.H. Bayraktar, E.F. Morgan, G.L. Niebur, G.E. Morris, E.K. Wong, T.M. Keaveny, Comparison of the elastic and yield properties of human femoral trabecular and cortical
bone tissue, J. Biomech. 37 (2004) 2735.
[41] T. Ozaki, H. Matsumoto, S. Watanabe, S. Hanada, Beta Ti alloys with low Young's
modulus, Mater. Trans. 45 (2004) 27762779.
[42] H.W. Jeong, Y.S. Yoo, Y.T. Lee, J.K. Park, Elastic softening behavior of TiNb single
crystal near martensitic transformation temperature, J. Appl. Phys. 108 (2010)
063515. http://dx.doi.org/10.1063/1.3486212.
[43] L. You, X. Song, A study of low Young's modulus TiNbZr alloys using d electrons alloy theory, Scr. Mater. 67 (1) (2012) 5760. http://dx.doi.org/10.1016/
j.scriptamat.2012.03.020.
[44] M. Tane, S. Akita, T. Nakano, K. Hagihara, Y. Umakoshi, M. Niinomi, H. Mori, H.
Nakajima, Low Young's modulus of TiNbTaZr alloys caused by softening in
shear moduli C and C44 near lower limit of body-centered cubic phase stability,
Acta Mater. 58 (2010) 67906798.
[45] M. Tane, S. Akita, T. Nakano, K. Hagihara, Y. Umakoshi, M. Niinomi, H. Nakajima,
Peculiar elastic behavior of TiNbTaZr single crystals, Acta Mater. 56 (2008)
28562863.
[46] J.F. Nye, Physical Properties of Crystals, Oxford Press, 1957..

Das könnte Ihnen auch gefallen