Sie sind auf Seite 1von 16

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

Available online at www.sciencedirect.com

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

Research paper

Porous titanium materials with entangled wire structure for


load-bearing biomedical applications

Guo He ∗ , Ping Liu, Qingbiao Tan


State Key Lab of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University,
Shanghai 200240, China

A R T I C L E I N F O A B S T R A C T

Article history: A kind of porous metal-entangled titanium wire material has been investigated in terms
Received 20 July 2011 of the pore structure (size and distribution), the strength, the elastic modulus, and the
Received in revised form mechanical behavior under uniaxial tensile loading. Its functions and potentials for surgical
30 September 2011 application have been explained. In particular, its advantages over competitors (e.g.,
Accepted 30 September 2011 conventional porous titanium) have been reviewed. In the study, a group of entangled
Published online 12 October 2011 titanium wire materials with non-woven structure were fabricated by using 12–180 MPa
forming pressure, which have porosity in a range of 48%–82%. The pores in the materials
Keywords: are irregular in shape, which have a nearly half-normal distribution in size range. The yield
Entangled titanium wire materials strength, ultimate tensile strength, and elastic modulus are 75 MPa, 108 MPa, and 1.05 GPa,
Porous titanium respectively, when its porosity is 44.7%. The mechanical properties decrease significantly as
Biomaterials the porosity increases. When the porosity is 57.9%, these values become 24 MPa, 47.5 MPa,
Implants and 0.33 GPa, respectively. The low elastic modulus is due to the structural flexibility of the
Elastic modulus entangled titanium wire materials. For practical reference, a group of detailed data of the
porous structure and the mechanical properties are reported. This kind of material is very
promising for implant applications because of their very good toughness, perfect flexibility,
high strength, adequate elastic modulus, and low cost.
⃝c 2011 Elsevier Ltd. All rights reserved.

1. Introduction fect. These advantages have facilitated the practical clini-


cal applications of titanium and its alloys, such as artificial
Metallic biomaterials are preferential choices in load-bearing titanium alloy joint (Long and Rack, 1998), dental titanium
applications in the field of bone tissue engineering and re- implant (Bréme et al., 1993), coronary titanium–nickel stent
generative medicine due to their outstanding mechanical (Eigler et al., 1993), etc. In the past decades, attempts were
strength and good toughness. Among them, titanium and its frequently made to seek optimal, mechanical and biocompat-
alloys exhibit excellent biocompatibility, good corrosion re- ible properties for titanium implants (Niinomi, 2008; Geetha
sistance, and relatively lower stiffness (compared with other et al., 2009), so as to remedy the mismatch of the elastic mod-
metallic biomaterials such as cobalt-based alloys, stainless uli (between titanium implant and cortical/cancellous bone)
steel, etc.), which are beneficial to the durability of the load- and the bioinert properties of the titanium implant surface.
bearing implants and reduction of the stress-shielding ef- The former can lead to stress-shielding over the surround-

∗ Corresponding author. Tel.: +86 21 3420 2643; fax: +86 21 3420 2643.
E-mail address: ghe@sjtu.edu.cn (G. He).
c 2011 Elsevier Ltd. All rights reserved.
1751-6161/$ - see front matter ⃝
doi:10.1016/j.jmbbm.2011.09.016
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 17

ing bone, which results in resorption of the bone and asep- (Banhart, 2001; Körner and Singer, 2000) such as ‘foaming by
tic loosening of the implant (Jacobs et al., 1993); the latter gas injection’ and ‘foaming with blowing agents’, which have
may cause poor bone–implant interfacial bonding, because been widely used in the aluminum foaming industry, are in-
the bulk titanium surface develops interfacial fibrous tissue adequate to produce porous titanium for orthopedic applica-
to form encapsulation that isolates the implants from their tions because only closed-cell structure can be obtained. The
surroundings (Chang et al., 1996). In order to solve these conventional metallurgical method by titanium powder sin-
problems, alloying is a metallurgical method to reduce the tering was used to fabricate porous titanium successfully (Oh
elastic modulus via forming full beta structure. So far the et al., 2003), but the porosity was very limited. When space-
elastic modulus as low as thirty GPa has been achieved in holding material was introduced in the powder sintering pro-
some beta titanium alloys (Geetha et al., 2009; Laheurte et al., cessing the porosity could been increased up to 80% or even
2010). But compared with cortical bone (4.4–28.8 GPa) and can- more (Banhart, 2001; Garrett et al., 2006; Wen et al., 2001), and
cellous bone (0.01–3.0 GPa) (Geetha et al., 2009), the beta ti- this method could produce bimodal pore sizes by controlling
tanium alloy is still stiffer. It seems that only a very limited the space-holder size and the titanium particle diameter. Fur-
room for reduction of the elastic modulus of the bulk titanium thermore, a modified powder metallurgical technique called
alloy remains. The surface modifications on titanium im- ‘sponge replication’ (Cachinho and Correia, 2008; Lee et al.,
plants have previously demonstrated that bioinert titanium 2009; Li et al., 2005) or ‘impregnating method’ (Zhao et al.,
metal could be converted into bioactive material through 2008) has been used in fabrication of the porous titanium
specific chemical and thermal treatments (Fujibayashi et al., and its alloys, by which the pore size, shape, and distribution
2001; Kim et al., 2000; Nishiguchi et al., 1999), by which the can be well-adjusted in a large scale. Some interconnected
smooth titanium surface was processed into a specific porous porous structures that perfectly mimic the spatial morphol-
structure (Fujibayashi et al., 2004; Kim et al., 2000). These pre- ogy of the cancellous bone can be easily reproduced by this
vious works implied the superiority of the porous titanium method. However, the major drawbacks of the above metal-
materials as implants. In fact, porous biomaterials have re-
lurgical methods are the difficulties to avoid contamination
ceived a lot of attention in past ten years. The porous ti-
and impurity phases in the titanium materials, which sig-
tanium can provide adequate macro/micro-pores for bone
nificantly deteriorate the mechanical properties of the tita-
ingrowth, vascularization, and flow transport of nutrients and
nium struts. In addition, the sintered titanium struts likely
metabolic waste (Manukyana et al., 2010), which is supposed
contain cracks, unbonded contact nodes between the powder
to be osteoconductive, osteoinductive and capable of osseoin-
particles, and other metallographic defects, thus, the porous
tegration (Albrektsson and Johansson, 2001; Stevens, 2008).
titanium materials cannot bear tensile stress and even ex-
It is well accepted so far that the titanium material
hibit ‘brittle’ nature. This is the reason that almost all the
with porous structure, which can imitate cancellous bone in
investigations did not report the tensile strength of the pow-
structure and properties, facilitates proliferation of cells into
der sintered porous titanium materials in the literature. This
the structure and provides space for bio-factor (cell, gene
limitation implies a risk of the orthopedic applications of the
and/or protein) delivery (Hollister, 2005). Such bone tissue
porous titanium in vivo where a tensile or impact load may
ingrowth through the pores forms stable long-term anchorage
be applied on the implant.
for biological fixation of the implant (Garrett et al., 2006;
In recent years a rapid prototyping technique has been uti-
Krishna et al., 2007). It is important that the porous network
lized to fabricate porous titanium by using laser power to sin-
must interconnect and the appropriate pore size for in vivo
ter/remelt titanium powders (Hollander et al., 2006; Krishna
bone ingrowth should be in the range of 100–500 µm (Bungo
et al., 2007, 2008; Mullen et al., 2008; Traini et al., 2008), which
et al., 2006; Freyman et al., 2001; Garrett et al., 2006; Holy
et al., 2000; Li et al., 2007). Some investigations validated is based on a CAD and layer-by-layer manufacturing process.
that the porous titanium implants could be improved in It is of great benefit to accurately controlling the pore shape,
osteoconductive properties as porosity increased, and the size and distribution by beforehand design and programming,
amount of new bone growth obviously increased as the so that an ideal porous architecture can be constructed. With
pore size increased (Bungo et al., 2006; Li et al., 2007). alternative fabrication approaches, this technique has been
Another benefit of porous titanium is that the effective elastic developed into ‘direct laser sintering’ (Hollander et al., 2006;
modulus can be tuned to match the modulus of bone to Krishna et al., 2007; Traini et al., 2008) and ‘selective laser
reduce the problems associated with stress-shielding. The melting’ (Krishna et al., 2008; Mullen et al., 2008). The rapid
mechanical properties of porous titanium are dependent on prototyping technique provides good feasibility and excellent
porosity, porous morphology and pore size distribution as flexibility for production of custom-built metallic implants
these determine the spatial shape and size of the struts or with complicated structures. However, the porous titanium
cell-walls (for honeycombed structure) between the pores production also suffers from impurities, inclusions and other
which are bearing the load. The nature of the struts or defects associated with the laser sintering or melting process-
cell-walls together with the holistic architecture constructed ing, which significantly deteriorate the mechanical properties
by the struts is crucial for the macro-mechanical behaviors such as tensile-stress resistance and toughness. In the view
of the bulk porous titanium. All these key macroscopic of physical metallurgy, the struts in the porous titanium with
parameters are controlled by the fabrication methods of the as-sintered (laser sintering) or as-solidified (laser melting) mi-
porous titanium materials. crostructure usually exhibits much lower ductility than that
There are many techniques to produce porous tita- of the wrought or plastically deformed counterparts. It is
nium, and various methods may lead to different porosities unlikely that the laser sintered/melted porous titanium im-
and morphologies. The earliest metal foaming techniques plants can stand against tensile and impact stress, though
18 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

there are very few data published (Hollander et al., 2006; Kr- 1996; Paquay et al., 1997). And other works attempted to mod-
ishna et al., 2007, 2008; Mullen et al., 2008; Traini et al., 2008). ify the mesh with bioactive coating (Vehof et al., 2000) and
Apart from laser rapid prototyping, a similar 3D deposition load with growth factor (Vehof et al., 2002) or bone marrow
approach was also used to fabricate porous titanium, in which stromal cells (Dolder et al., 2003). They all had demonstrated
the deposited 3D porous structure was strengthened by dry- the good osteoconductive properties of porous titanium fiber
ing in air and sintering in high vacuum (Li et al., 2006). This mesh. However, few works deal with the mechanical behav-
approach does not need laser source and control system in- iors of the titanium fiber mesh. The relationship between the
volved. Despite the fact that a good 3D mesh structure was mesh structure and the mechanical properties has not been
constructed, its low tensile strength and low ductility can be established. Furthermore, the basic aspects in terms of the
expected due to the similar above reasons. fabrication techniques (especially for the 3D porous titanium
The mechanical properties of the porous metallic wire materials), the structure characterization, and the me-
materials are closely related to the porosity and the porous chanical properties tailoring, are still open for investigation.
architecture (including pore size, shape, and distribution). A data base including all above essentials is urgently needed,
The relationship of the mechanical properties and the because it is necessary for the bio-applications. In this study,
porosity has already been established empirically which can we report a group of data on the structure and mechanical
be expressed as the well-known Gibson–Ashby equation properties of the 3D ‘entangled titanium wire material’. Spe-
(Gibson and Ashby, 1997): cial emphasis is put on the tensile behavior and the rela-
 ∗ n 1 tionship between the porosity and the mechanical properties.
E∗ ρ This work will provide researchers and surgeons/clinicians
= C1 (1)
Es ρs with the valuable data of such tough porous titanium materi-
 ∗ n2
σ∗ ρ als for the load-bearing orthopedic applications.
= C2 (2)
σs ρs
where E, σ and ρ denote the elastic modulus, strength and
2. Materials and methods
density, respectively. The superscript ‘∗’ indicates the value of
the porous material, and the subscript ‘s’ indicates the value
2.1. Fabrication of the entangled titanium wire materials
of the dense material. C1 and C2 are constants related to ma-
terial and experimental conditions. n1 and n2 are exponen-
The mechanical behavior of the porous titanium wire materi-
tials related to porous structure. It is clear that both strength
als strongly depends on raw material (titanium composition,
and elastic modulus of the porous materials decrease as the
treatment status, and wire diameter), entangled wire archi-
porosity increases. A contradiction is that low elastic modulus
tecture, and porosity. Our work (Liu, 2010) demonstrated that
(i.e., comparable to cortical bone) corresponds to low strength.
a smaller diameter yields higher strength at the same poros-
Considering the defects and contaminations in the struts, the
ity. In this study, commercial pure titanium (Grade 1) wires
porous titanium with low elastic modulus is too weak to be with diameters of 0.08 and 0.15 mm, respectively, are selected
used in load-bearing implants. To enhance its strength, espe- as raw materials. The impurities in the titanium wires are
cially for resistance to tensile stress and impact load, will be chemically determined to be N: 0.02%, O: 0.15%, H: 0.015%, C:
a challenge in the real orthopedic application. 0.008% (all in weight percent). Their Young’s modulus, yield
In order to improve the toughness and tensile strength strength, and ultimate strength are 116 GPa, 300 MPa, and
of the porous titanium, a novel porous titanium material so 370 MPa, respectively. The spatial architecture constructed
called ‘entangled wire material’ (Liu et al., 2009, 2010; Tan with wires is controlled by the wire-twisting methods and
et al., 2009) has been developed recently in our group. Its intertwisting procedures. Two styles (non-woven and quasi-
prominent advantage is that the strong metal wires acting ordered) have been developed in our lab which show signifi-
as struts construct the spatial porous structure, so that the cant difference in mechanical behavior between the two types
drawbacks of the as-sintered/as-solidified struts (e.g., metal- of the architecture. Their fabrication methods have been de-
lurgical defects, contaminations, etc.) can be completely over- scribed in the literature (Liu et al., 2009, 2010; Tan et al.,
come. Compared with the ‘brittle’ nature of the conventional 2009). Generally, the porous titanium wire materials with
porous titanium, the entangled titanium wire porous material non-woven architecture exhibit higher strength and higher
is very tough and exhibits excellent superelasticity. Its supe- stiffness than that with quasi-ordered architecture. In this
riority in mechanical properties supposes the potentials for study, we focus on the non-woven titanium wire materials,
orthopedic applications such as hip joint implant, lumber in- which are composed of entangled wires, wire cross points,
terbody fusion device, dental implant, etc. and pores between the wires. The shape, size, and morphol-
It should be noticed that the ‘entangled wire material’ is ogy of the pores in the material are determined by the spatial
a kind of fiber (wire) mesh, in which the entangled wire can architecture of the entangled wires, which is strongly decisive
be either stochastic (like non-woven) (Liu et al., 2009, 2010) for the mechanical behaviors and properties.
or ordered (Liu et al., 2010; Tan et al., 2009). Originally, ti- A simple intertwisting procedure following forming in
tanium wire (fiber) materials were introduced for producing a special mold can be used to prepare the non-woven
porous-surfaced implants (Ducheyne et al., 1974). The porous entangled titanium wire material. The spatial architecture of
titanium fiber layer bonded on the surface of the implant pro- the entangled wires can be constructed via various winding
vides space for bone ingrowth to fix the implant. In order to and forming methods. A general procedure for forming non-
examine its validity, some works focused on the bone forma- woven titanium wire materials is as follows: firstly the wire
tion and wound healing in titanium fiber mesh (Chang et al., is entangled by winding and twisting to form a precursory
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 19

bundle of twisted wires. Then the bundle is put into a determined by means of mathematical statistics of the
specially designed steel mold for compaction and shaping. measurements. Five viewpoints are chosen from the cross
A high pressure in a range of 12–180 MPa is necessary. and longitudinal sections of each sample.
Generally, the porous materials with larger porosity need The uniaxial tensile stress–strain behavior of the porous
lower forming pressure, but smaller porosity need higher titanium wire materials has been performed by using
forming pressure. The holding time under certain pressure Shimadzu AG-100KNA testing machine at room temperature.
and at high temperature (around 650 ◦ C) is about 10 min for The tests are conducted under displacement control, with a
avoiding wire-rebound and shape-finalizing. At this stage, the crosshead speed of 0.05 mm/s, which corresponds to a mean
wires cross points in the material are mobile, so that the bulk strain rate of about 2.27 × 10−3 s−1 . The stress is calculated
material exhibits flexible properties but low yield strength by dividing the tensile force by the initial cross sectional
and low stiffness. For improving its strength a sintering area of the sample, and the strain is the ratio between the
processing is needed which can be performed in a furnace sample elongation and the initial gauge length. The elastic
in argon atmosphere. The sintering temperature is about modulus is determined from the initial near-elastic part of the
1200 ◦ C and holding time is 60 min. After the sintering a stress–strain curve. The deformation and failure mechanism
step-cooling followed with annealing at 650 ◦ C is carried out of the porous titanium wire materials are investigated by
for plasticity recovering of the titanium wires. The optimized using the SEM.
sintering process has been described in detail elsewhere (Liu
et al., 2010). After sintering some wire cross points in the
material are fixed by forming sintered joints, so that the yield 3. Results
strength and stiffness can be significantly improved.
So far, several shapes such as cylinders, rectangular 3.1. The porous structure in the entangled titanium wire
prisms, tubes, sheets, and flat dog-bones have been made in materials
our lab. The cylindrical samples with 10 mm in diameter and
20 mm in length are for compressive test. The flat dog-bones A macroscopical 3D image for the entangled titanium wire
with the gauge length of about 22 mm and the section size of material is shown in Fig. 1(a). By means of the tomography
5 mm × 3 mm in the gauge are fabricated for tensile test. reconstruction whole the sample can be revealed along any
cross section. As an example, Fig. 1(b) displays a section view
2.2. Characterization of the entangled titanium wire which clearly shows the inner macrostructure of the non-
materials woven entangled titanium wire material. Although the pores
in the material are approximately homogeneous, some of the
The spatial architecture of the porous titanium wire materials titanium wires tend to assemble into wire bundles during
can be revealed by means of X-ray tomography which the forming process and preferentially turn toward specific
allows visualizing the inner volume of the material without orientation, e.g., along longitudinal axis of the sample. Thus,
sectioning or dismantling it. For quantitative analysis the spatial architecture is usually inhomogeneous in the
the 3D architecture can be mathematically reconstructed non-woven entangled titanium wire material. Fig. 2 shows
from 2D projection images collected by the tomography a tube-like sample and a group of section views across the
device while the material is rotated around a defined axis longitudinal axis via the tomography. The pore distribution in
(Garcia and Kardjilov, 2005). Fig. 1 shows a reconstructed the material is roughly homogeneous, but the arrangement of
3D porous material (left) and an arbitrary section (right). the entangled wires is inhomogeneous. This may lead to the
The homogeneity and the inner porous structure can be dispersion in the values of the mechanical properties.
statistically evaluated from the section view. For clearly In order to distinguish the entangled wire architecture
showing the local porous structure and the wire joints a and the pore structure, a close look was taken at the
scanning electron microscopy (SEM) JEOL-JSM6460 is used. microstructure by using SEM. Fig. 3(a) and (d) show the
The porosity of the porous titanium wire materials can be surface images of the entangled wire structures made of two
directly obtained by mass–volume calculation method with different diameter wires, and Fig. 3(b) and (e) show their
the following formula: section images. It is clear that the wire diameter influences
the pore size. Larger wire diameter leads to larger pores
M
 
P= 1− × 100% (3) (compared (a) with (d), (b) with (e)) although both materials
Vρs are in very similar porosity. The sample made of 0.08 mm-
where M is weight of the sample; V is volume of the sample; diameter wire with 56% porosity has about 60% pores smaller
ρs is the density of the titanium wire which is 4.51 g/cm3 . than 100 µm and about 40% pores in range of 100–300 µm as
Since all the pores in the material are open and connected, shown in Fig. 3(c). The other made of 0.15 mm-diameter wire
the calculated porosity is very close to the measurement with with 56.6% porosity has about 45% pores smaller than 100 µm
Archimedes method (Liu et al., 2010). and about 55% pores larger than 100 µm. The largest pores are
In order to measure the pore size quantitatively, an above 600 µm as shown in Fig. 3(f). It is undoubted that the
equivalent pore diameter is defined as the interspace between entangled titanium wire materials with larger porosity can
twisted wires. The measurements can be performed under provide more pores for bone ingrowth and vascularization.
optical microscope by using the commercial Image Pro The materials with smaller porosity can only provide a few
Discovery software (the spatial pores are taken as two- pores large enough for the implants according to Fig. 4 and
dimensional pores). Thus, the pore size distributions are Table 1. Even though there are, so far, no data on how many
20

Table 1 – Data of the porous structure for the entangled Ti wire materials (P1–P4: 0.08 mm wire, P3# : 0.15 mm wire).

No. Forming Green compact Sintering and heat treatment After sintering and heat treatment
pressure (MPa) porosity (%) parameters
Open porosity Total porosity Maximum Average Pore size
(%) (%) pore pore distribution
size (µm) size (µm) (µm)
P1-1 12.6 ± 0.8 82.0 ± 0.2 1100 ◦ C × 30 min + 650 ◦ C × 30 min 80.9 ± 0.1 81.2 ± 0.1 673 180 50–450
P1-2 12.6 ± 0.8 82.0 ± 0.2 1100 ◦ C × 60 min + 650 ◦ C × 30 min 80.4 ± 0.1 80.7 ± 0.1 656 176 50–450
P1-3 12.6 ± 0.8 82.0 ± 0.2 1100 ◦ C × 90 min + 650 ◦ C × 30 min 79.8 ± 0.1 80.1 ± 0.1 639 171 50–450
P1-4 12.6 ± 0.8 82.0 ± 0.2 1200 ◦ C × 30 min + 650 ◦ C × 30 min 79.2 ± 0.1 79.6 ± 0.1 623 167 45–425
P1-5 12.6 ± 0.8 82.0 ± 0.2 1200 ◦ C × 60 min + 650 ◦ C × 30 min 78.7 ± 0.1 79.1 ± 0.1 606 163 45–425
P1-6 12.6 ± 0.8 82.0 ± 0.2 1200 ◦ C × 90 min + 650 ◦ C × 30 min 78.1 ± 0.1 78.5 ± 0.1 589 159 45–400
P1-7 12.6 ± 0.8 82.0 ± 0.2 1300 ◦ C × 30 min + 650 ◦ C × 30 min 77.6 ± 0.1 78.0 ± 0.1 573 155 45–400
P1-8 12.6 ± 0.8 82.0 ± 0.2 1300 ◦ C × 60 min + 650 ◦ C × 30 min 77.0 ± 0.1 77.4 ± 0.1 556 151 45–375
P1-9 12.6 ± 0.8 82.0 ± 0.2 1300 ◦ C × 90 min + 650 ◦ C × 30 min 76.4 ± 0.1 76.8 ± 0.1 539 146 45–375
P2-1 35.9 ± 1.0 70.0 ± 0.2 1100 ◦ C × 30 min + 650 ◦ C × 30 min 68.8 ± 0.1 69.3 ± 0.1 523 142 40–350
P2-2 35.9 ± 1.0 70.0 ± 0.2 1100 ◦ C × 60 min + 650 ◦ C × 30 min 68.3 ± 0.1 68.8 ± 0.1 506 138 40–350
P2-3 35.9 ± 1.0 70.0 ± 0.2 1100 ◦ C × 90 min + 650 ◦ C × 30 min 67.8 ± 0.1 68.3 ± 0.1 489 134 40–350
P2-4 35.9 ± 1.0 70.0 ± 0.2 1200 ◦ C × 30 min + 650 ◦ C × 30 min 67.3 ± 0.1 67.8 ± 0.1 473 131 35–325
P2-5 35.9 ± 1.0 70.0 ± 0.2 1200 ◦ C × 60 min + 650 ◦ C × 30 min 66.8 ± 0.1 67.3 ± 0.1 457 128 35–325
P2-6 35.9 ± 1.0 70.0 ± 0.2 1200 ◦ C × 90 min + 650 ◦ C × 30 min 66.2 ± 0.1 66.7 ± 0.1 440 124 35–300
P2-7 35.9 ± 1.0 70.0 ± 0.2 1300 ◦ C × 30 min + 650 ◦ C × 30 min 65.7 ± 0.1 66.2 ± 0.1 424 121 35–300
P2-8 35.9 ± 1.0 70.0 ± 0.2 1300 ◦ C × 60 min + 650 ◦ C × 30 min 65.2 ± 0.1 65.7 ± 0.1 408 118 35–275
P2-9 35.9 ± 1.0 70.0 ± 0.2 1300 ◦ C × 90 min + 650 ◦ C × 30 min 64.6 ± 0.1 65.1 ± 0.1 391 115 35–275
P3-1 81.5 ± 1.2 59.0 ± 0.2 1100 ◦ C × 30 min + 650 ◦ C × 30 min 57.8 ± 0.1 58.4 ± 0.1 375 112 30–250
P3-2 81.5 ± 1.2 59.0 ± 0.2 1100 ◦ C × 60 min + 650 ◦ C × 30 min 57.3 ± 0.1 57.9 ± 0.1 358 109 30–250
P3-3 81.5 ± 1.2 59.0 ± 0.2 1100 ◦ C × 90 min + 650 ◦ C × 30 min 56.8 ± 0.1 57.4 ± 0.1 341 106 30–250
P3-4 81.5 ± 1.2 59.0 ± 0.2 1200 ◦ C × 30 min + 650 ◦ C × 30 min 56.4 ± 0.1 57.0 ± 0.1 325 103 25–225
P3-5 81.5 ± 1.2 59.0 ± 0.2 1200 ◦ C × 60 min + 650 ◦ C × 30 min 55.9 ± 0.1 56.5 ± 0.1 308 100 25–225
P3-6 81.5 ± 1.2 59.0 ± 0.2 1200 ◦ C × 90 min + 650 ◦ C × 30 min 55.4 ± 0.1 56.0 ± 0.1 291 96 25–200
P3-7 81.5 ± 1.2 59.0 ± 0.2 1300 ◦ C × 30 min + 650 ◦ C × 30 min 54.9 ± 0.1 55.5 ± 0.1 275 93 25–200
P3-8 81.5 ± 1.2 59.0 ± 0.2 1300 ◦ C × 60 min + 650 ◦ C × 30 min 54.4 ± 0.1 55.0 ± 0.1 258 90 25–175
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S

P3-9 81.5 ± 1.2 59.0 ± 0.2 1300 ◦ C × 90 min + 650 ◦ C × 30 min 53.9 ± 0.1 54.5 ± 0.1 241 87 25–175
P4-1 180.4 ± 1.5 48.0 ± 0.2 1100 ◦ C × 30 min + 650 ◦ C × 30 min 46.8 ± 0.1 47.5 ± 0.1 224 84 20–150
P4-2 180.4 ± 1.5 48.0 ± 0.2 1100 ◦ C × 60 min + 650 ◦ C × 30 min 46.4 ± 0.1 47.1 ± 0.1 206 81 20–150
P4-3 180.4 ± 1.5 48.0 ± 0.2 1100 ◦ C × 90 min + 650 ◦ C × 30 min 46.0 ± 0.1 46.7 ± 0.1 194 78 20–150
P4-4 180.4 ± 1.5 48.0 ± 0.2 1200 ◦ C × 30 min + 650 ◦ C × 30 min 45.6 ± 0.1 46.3 ± 0.1 182 75 15–135
P4-5 180.4 ± 1.5 48.0 ± 0.2 1200 ◦ C × 60 min + 650 ◦ C × 30 min 45.2 ± 0.1 45.9 ± 0.1 171 72 15–135
P4-6 180.4 ± 1.5 48.0 ± 0.2 1200 ◦ C × 90 min + 650 ◦ C × 30 min 44.8 ± 0.1 45.5 ± 0.1 159 69 15–135
5 (2012) 16–31

P4-7 180.4 ± 1.5 48.0 ± 0.2 1300 ◦ C × 30 min + 650 ◦ C × 30 min 44.4 ± 0.1 45.1 ± 0.1 148 67 15–120
P4-8 180.4 ± 1.5 48.0 ± 0.2 1300 ◦ C × 60 min + 650 ◦ C × 30 min 44.0 ± 0.1 44.7 ± 0.1 136 64 15–120
P4-9 180.4 ± 1.5 48.0 ± 0.2 1300 ◦ C × 90 min + 650 ◦ C × 30 min 43.5 ± 0.1 44.2 ± 0.1 125 61 15–120
P3# -1 81.5 ± 1.2 59.0 ± 0.2 1100 ◦ C × 90 min + 650 ◦ C × 30 min 57.3 ± 0.1 57.9 ± 0.1 686 131 35–400
P3# -2 81.5 ± 1.2 59.0 ± 0.2 1200 ◦ C × 90 min + 650 ◦ C × 30 min 56.0 ± 0.1 56.6 ± 0.1 645 122 30–350
P3# -3 81.5 ± 1.2 59.0 ± 0.2 1300 ◦ C × 90 min + 650 ◦ C × 30 min 54.5 ± 0.1 55.1 ± 0.1 605 116 30–300
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 21

Fig. 1 – Tomography images of the entangled Ti wire materials with non-woven structure. (a) The reconstructed 3D image
of a rectangular prism-like sample, and (b) the sectioned 3D image of this sample.

Fig. 2 – Tomography image of a tube-like entangled Ti wire material. The section views show the porous structure and its
homogeneity.

pores are necessary for the implant for forming effective titanium wire materials with various porosities are provided
fixing with the surrounding tissue, it is well accepted that in Table 1, including the forming pressure which is needed
the pore sizes in range of 100–500 µm are favorable for the for the specific porosity, and the sintering and heat treatment
surgical applications (Bungo et al., 2006; Freyman et al., 2001; parameters. The green compact porosity in the table was
Garrett et al., 2006; Holy et al., 2000; Li et al., 2007). But calculated according to formula (3). The open porosity and
how these materials work in vivo needs further biological the total porosity were evaluated after sintering and heat
experiments. Despite all that, one can be assured that the treatment by using the hydrostatic weighing method (Liu
0.15 mm-diameter titanium wire results in larger pores than et al., 2010; Taylor et al., 1999).
0.08 mm-diameter wire (Fig. 3), thus is preferential when the
porosity is the same. 3.2. Tensile behavior and property of the entangled
To establish the relationship between the pore size and titanium wire materials
the porosity, a group of samples made of 0.08 mm-diameter
titanium wire with different porosities were evaluated by Under uniaxial tensile stress the entangled titanium wire
measuring the pore size. When the porosity is 48%, the pore materials behave partially like a rubber or biopolymer with
size is relatively homogeneous and more than 70% pores non-linearity, high elongation at break, and low elastic
are under 100 µm as shown in Fig. 4(a). For about 60% modulus, because the entangled wire structure is similar to
porosity, near 50% pores are larger than 100 µm. A larger the molecular networks in rubber or polymer. The typical
porosity corresponds to a larger percent of the pores above stress–strain curves are composed of initial near-elastic stage
100 µm (Fig. 4). The sample with 82% porosity has about (as indicated by OA in Fig. 6), the knee of the curve (A in Fig. 6),
25% pores under 100 µm and 75% pores above 100 µm as plastic deformation stage (AB in Fig. 6), and failure start point
shown in Fig. 4(d). A distinct relationship between the pore (B in Fig. 6). The elastic modulus is defined as the slope of
size and the porosity has been shown in Fig. 5. For reference, the initial OA stage. The yield strength is determined from
all the data of the porous structure for the entangled A point. It is easily understandable that the OA corresponds
22 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

Fig. 3 – Entangled wire architectures after sintering and the pore size distributions. (a)–(c): for wire diameter 0.08 mm; and
(d)–(f): for wire diameter 0.15 mm.

Fig. 4 – Pore size distributions in the entangled Ti wire materials with various porosities (wire diameter 0.08 mm).
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 23

Fig. 7 – Effect of the porosity on the tensile strength and


the elastic modulus.
Fig. 5 – Dependence of the pore size on the porosity.

Fig. 8 – Effect of the Ti wire diameter on the tensile


behavior of the entangled Ti wire materials.

Fig. 6 – Nominal tensile stress–strain curves of the


entangled Ti wire materials with various porosities.
strength of 47.5–108 MPa can imitate that of the natural
cortical bone (ultimate tensile strength: 9.9–151 MPa) (Donald
to the structure elastic deformation, but AB corresponds to et al., 1974). The elastic modulus in range of 0.04–1.2 GPa is
the structure plastic deformation via wires sliding each other. comparable to cancellous bone (0.01–3.0 GPa), but is too low
If the material is in high density (small porosity) the sliding to compare with the cortical bone (4.4–28.8 GPa) (Geetha et al.,
can be blocked, and then the structure plastic deformation 2009). If they are used for replacement of the cortical bone,
becomes difficult. Such effect has been clearly revealed in strengthening must be needed. Another contradiction is that
Fig. 6 where the yield point moves to the failure start point the lower porosity corresponds to larger elastic modulus and
as the porosity decreases. For the sample with 44.7% porosity, higher strength, but smaller pore size which is unfavorable
it behaves like a brittle alloy as shown in Fig. 6 for which there for the bone ingrowth. Further modification on the porous
is no distinct plastic deformation stage. structure is desirable.
It should be noted that the entangled titanium wire The titanium wire diameters also influence the tensile
materials exhibit very low stiffness due to the pliable behaviors and properties as shown in Fig. 8. With the similar
spatial architecture. The entangled wire structure with porosity, the material made of 0.08 mm-diameter titanium
higher density (i.e. smaller porosity) should be more stable wire exhibits higher strength and larger elastic modulus
under loading than that with lower density (larger porosity). but smaller elongation than that of 0.15 mm-diameter
Therefore, the material with lower porosity has higher titanium wire. For reference, all the tensile properties of the
strength and larger elastic modulus. For example, 44.7% entangled titanium wire materials with various porosities are
porosity corresponds to 75 MPa yield strength, 108 MPa summarized in Table 2. The sample numbers correspond to
ultimate tensile strength, and 1.05 GPa elastic modulus. For that in Table 1.
57.9% porosity, these values are only 24 MPa, 47.5 MPa, and In addition, good toughness of these materials is expected
0.33 GPa, respectively. All the three values reveal a significant because they exhibit a typical ductile manner with large
dependence on the porosity, which can be plotted in Fig. 7. strain to failure under tensile loading. In order to evaluate
The yield strength of 24–75 MPa and the ultimate tensile their toughness values, one can measure the area underneath
24 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

Table 2 – Tensile properties of the entangled Ti wire materials (P1–P4: 0.08 mm wire, P3∗ : without sintering; P3# :
0.15 mm wire).

No. Total porosity Yield strength UTS (MPa) Elastic modulus Elongation at F-max
(%) (MPa) (MPa) (%)
P1-1 81.2 ± 0.1 2.6 ± 0.1 5.7 ± 0.3 41.1 ± 0.6 54.7 ± 0.4
P1-2 80.7 ± 0.1 3.1 ± 0.1 6.8 ± 0.3 46.0 ± 0.5 53.4 ± 0.3
P1-3 80.1 ± 0.1 3.8 ± 0.1 8.0 ± 0.1 51.4 ± 0.3 51.9 ± 0.1
P1-4 79.6 ± 0.1 4.1 ± 0.1 8.6 ± 0.1 54.1 ± 0.3 51.2 ± 0.1
P1-5 79.1 ± 0.1 4.4 ± 0.1 9.2 ± 0.1 56.8 ± 0.4 50.5 ± 0.1
P1-6 78.5 ± 0.1 4.8 ± 0.1 9.9 ± 0.1 60.0 ± 0.4 49.7 ± 0.1
P1-7 78.0 ± 0.1 5.1 ± 0.1 10.5 ± 0.1 62.7 ± 0.5 49.0 ± 0.2
P1-8 77.4 ± 0.1 5.6 ± 0.2 11.5 ± 0.2 67.2 ± 0.6 47.8 ± 0.3
P1-9 76.8 ± 0.1 7.0 ± 0.3 12.7 ± 0.3 72.6 ± 0.9 46.3 ± 0.4
P2-1 69.3 ± 0.1 8.8 ± 0.2 19.0 ± 0.4 114.6 ± 0.8 44.2 ± 0.3
P2-2 68.8 ± 0.1 9.6 ± 0.2 20.7 ± 0.4 122.0 ± 0.8 43.2 ± 0.2
P2-3 68.3 ± 0.1 10.5 ± 0.2 22.8 ± 0.3 131.3 ± 0.9 41.9 ± 0.2
P2-4 67.8 ± 0.1 11.3 ± 0.1 24.7 ± 0.1 139.7 ± 0.6 40.8 ± 0.1
P2-5 67.3 ± 0.1 11.7 ± 0.1 25.5 ± 0.1 143.4 ± 0.6 40.2 ± 0.1
P2-6 66.7 ± 0.1 12.5 ± 0.1 27.4 ± 0.3 151.8 ± 0.8 39.1 ± 0.1
P2-7 66.2 ± 0.1 13.1 ± 0.1 28.8 ± 0.4 158.0 ± 1.0 38.3 ± 0.2
P2-8 65.7 ± 0.1 14.3 ± 0.3 31.4 ± 0.4 169.5 ± 1.6 36.7 ± 0.3
P2-9 65.1 ± 0.1 18.2 ± 0.5 32.8 ± 0.5 184.5 ± 3.5 35.9 ± 0.2
P3∗ 59.0 ± 0.2 18.2 ± 0.5 35.8 ± 0.9 305.5 ± 4.4 10.6 ± 0.3
P3-1 58.4 ± 0.1 23.0 ± 0.2 44.9 ± 0.4 316.1 ± 1.8 33.5 ± 0.2
P3-2 57.9 ± 0.1 24.1 ± 0.2 47.5 ± 0.6 334.7 ± 1.2 32.6 ± 0.1
P3-3 57.4 ± 0.1 25.5 ± 0.1 50.6 ± 0.3 356.9 ± 1.1 31.6 ± 0.1
P3-4 57.0 ± 0.1 26.4 ± 0.2 52.7 ± 0.3 371.7 ± 1.1 30.9 ± 0.1
P3-5 56.5 ± 0.1 27.6 ± 0.3 55.5 ± 0.5 391.8 ± 1.4 30.0 ± 0.2
P3-6 56.0 ± 0.1 29.2 ± 0.2 59.1 ± 0.4 417.6 ± 1.3 28.8 ± 0.1
P3-7 55.5 ± 0.1 30.3 ± 0.3 61.5 ± 0.5 434.8 ± 1.4 28.0 ± 0.1
P3-8 55.0 ± 0.1 32.0 ± 0.4 65.3 ± 0.5 462.0 ± 2.2 26.7 ± 0.3
P3-9 54.5 ± 0.1 33.8 ± 0.5 69.4 ± 0.6 491.3 ± 4.3 25.3 ± 0.5
P4-1 47.5 ± 0.1 52.6 ± 0.5 80.9 ± 0.4 609.3 ± 4.6 22.5 ± 0.1
P4-2 47.1 ± 0.1 54.9 ± 0.2 83.8 ± 0.5 656.8 ± 3.8 21.9 ± 0.1
P4-3 46.7 ± 0.1 57.8 ± 0.2 87.7 ± 0.4 720.6 ± 3.5 21.0 ± 0.1
P4-4 46.3 ± 0.1 59.8 ± 0.3 90.3 ± 0.5 763.1 ± 3.4 20.4 ± 0.1
P4-5 45.9 ± 0.1 62.7 ± 0.4 94.1 ± 0.6 825.3 ± 4.2 19.5 ± 0.2
P4-6 45.5 ± 0.1 66.2 ± 0.4 98.7 ± 0.6 900.6 ± 4.0 18.5 ± 0.1
P4-7 45.1 ± 0.1 69.0 ± 0.5 102.4 ± 0.7 961.1 ± 4.6 17.7 ± 0.2
P4-8 44.7 ± 0.1 74.9 ± 0.5 108.1 ± 0.7 1054.4 ± 5.1 16.4 ± 0.4
P4-9 44.2 ± 0.1 77.3 ± 0.7 116.8 ± 0.5 1222.6 ± 4.3 14.4 ± 0.3
P3# -1 57.9 ± 0.1 16.6 ± 0.3 46.2 ± 0.4 144.9 ± 5.5 49.1 ± 0.5
P3# -2 56.6 ± 0.1 20.7 ± 0.5 50.9 ± 0.5 243.8 ± 4.9 46.5 ± 0.4
P3# -3 55.1 ± 0.1 25.0 ± 0.6 56.4 ± 0.7 347.6 ± 5.8 43.5 ± 0.3

the tensile stress–strain curves (Fig. 6) by taking the integral. 1974). The present materials exhibit 15%–40% tensile strain
The toughness value can be determined mathematically by: to failure as shown in Fig. 6.
 ε
energy f
= σdε (4)
volume 0 3.3. Failure of the entangled titanium wire materials
where ε is strain, εf is the strain upon failure, and σ is under tensile stress
stress. Then, the toughness values of the entangled titanium
wire materials (porosities in the range of 44.7%–57.8% as The deformation of the entangled titanium wire materials
indicated in Fig. 6) are in the range of 1.19–1.56 kJ/cm3 . under tensile loading initiates from the structure stretching
They are far above that of the titanium foams, of which and progresses accompanying local wires unwinding toward
the impact toughness is 2.4 kJ/m2 for 60% porosity and the tensile load. How many wires unwind and to what
1.3 kJ/m2 for 70% porosity (Kashef et al., 2010). Although there extent the unwinding attains depend on the porosity. For the
is limited literature on the toughness of porous titanium, it material made of 0.08 mm-diameter titanium wire with 57.9%
is undoubted that the toughness of the present materials is porosity, the sample behaves like a negative Poisson’s ratio
superior to that of many other porous titanium materials. material (Lakes, 1993) with auxetic behavior in the process
If compared with natural bone, the entangled titanium wire of unwinding (corresponding to AB stage as shown in Fig. 6).
material is obviously advantageous in toughness because the The auxeticity proceeds as the tension persists until most
bone is almost brittle under tensile loading and has small of the wires are stretched toward the tensile load as shown
compressive strain to failure (less than 2%) (Donald et al., in Fig. 9. Under the ultimate tensile stress the stretched
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 25

Fig. 9 – Deformation and failure during tension for the entangled Ti wire materials (wire diameter 0.08 mm,
porosity = 57.9%).

Fig. 10 – Fractographies of the entangled Ti wire materials (wire diameter 0.08 mm). (a) Failed samples with different
porosity after tension test, (b) and (c): magnified images of the zone of fracture for the sample with 45.9% porosity.

single wire fracture occurs and leads to rapid rupture of load and the compact entangled wire structure is completely
other neighboring wires due to the stress concentration. For destroyed.
the material with 45.9% porosity, the wires’ unwinding is
difficult because of the higher density (lower porosity). The
4. Discussion
entangled wire structure is remained in a high density until
failure occurs; as a result, the fracture is localized in a small 4.1. Mechanical property of the entangled titanium wire
zone as shown in Fig. 10(a)–(c). As the porosity increases the materials
fracture zone becomes indistinct. For the material with 78.5%
porosity, the sample becomes loosening after tension (Fig. 10), The entangled titanium wire material is a kind of porous
indicating that most of the wires unwind toward the tensile metal. Like other porous materials, their mechanical
26 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

properties strongly depend on the porosity. For conventional


open-cell foam a well-known relationship between the
mechanical properties and the porosity has been established
based on the initial linear elastic feature of the material by
using a simple procedure of structural mechanics (Gibson and
Ashby, 1997):
 ∗ 2
E∗ ρ
≈ (5)
Es ρs
 ∗ 3
σ∗ ρ 2
= 0.3 . (6)
σs ρs
These relationships have been validated in many open-cell
porous materials including polymer foams, sponge rubbers,
porous glasses, and porous metals (Gibson and Ashby,
1997). When we compare the mechanical properties of the
entangled titanium wire materials with the values predicted
by the above general formulas, distinct variances are found
between the measurements and the predictions as shown in
Fig. 11. At the side of small relative density (e.g., around 0.2),
the strengths are far below the predictions of Eq. (6), while
at the side of large relative density (e.g., around 0.55), the
strengths are far above the predictions (Fig. 11(a)). In addition,
a significant deviation can be found in the relative elastic
modulus as shown in Fig. 11(b). The measurements are about
two orders of magnitude less than the predictions. These
distinctions suggest very different deformation mechanism
and different mechanical behavior of the entangled titanium
wire materials compared to the open-cell foams.
The Gibson–Ashby equations were originally derived from
a simple stress relation considering an open-cell structure Fig. 11 – Relationship between the relative
under pressure as shown in Fig. 12(a). These relations were strength/modulus and the relative density for the
also considered being valid in the linear elastic stage for entangled Ti wire materials.
many porous materials under tension (Gibson and Ashby,
1997). In the nonlinear elastic stage, Warren and Kraynik
(1991) established a modified relation considering strain value
and the ‘strut’ rotation for the uniaxial tension as shown in
Fig. 12(b):
 ∗ 2  ∗ 2
σ∗ ρ ρ ρ
 ∗
= 1.1 ε + 3.74 ε2 + 0.0343 ε3 . (7)
Es ρs ρs ρs
For the entangled wire material, the structure can be
simplified as a 2D cell which is composed of wire segments
along various orientations and cross points as shown in
Fig. 12(c). The wire segments are usually curved and the
cross points may be free or fixed depending on the sintering
processing. The elastic deformation modes under low stress
can be the wire segments’ deflexion and unwinding if all Fig. 12 – Simplified structure models of the porous
the points are fixed. In this case the material should be in materials. (a) Linear elastic structure model of open-cell
relatively high stiffness and high yield strength. If all the foams under compression (Gibson and Ashby, 1997), (b)
cross points are free (without sintering) the structure should Nonlinear elastic structure model of open-cell foams under
be very flexible because the sliding can be triggered under tension (Warren and Kraynik, 1991), (c) Friction-slippage
very low stress. In such a case the material exhibits very structure model of the entangled Ti wire materials under
low stiffness and low yield strength. In as-sintered state, the tension.
free cross points and the sintered joints probably coexist in
the material. Although it is difficult to count the percentages
of the joints and the free cross points, one can affirm that materials have higher local contacted stress which is helpful
the joints are dominant at low porosity and the free cross for forming joints during sintering. Otherwise, the materials
points are in majority at high porosity. This can be easily with high porosity (loose structure) are difficult to form joints
understood because the wire cross points in the compact during sintering because of small local contacted stress in
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 27

the cross points. It is undoubted that the joints contribute to


the structure stability and enhance the stiffness and the yield
strength.
Considering 2D situation, the entangled wire materials
would behave similar to the open-cell foams if the entire cross
points became ideal joints and the joints were strong enough
and did not break before the material failure. In this case both
the 2D structure models (Fig. 12(b) and (c)) were similar, and
the mechanical properties of the entangled wire materials
became comparable to the open-cell foams. Considering 3D
situation, the materials with such strong joints may behave
similar to the truss material (Wallach and Gibson, 2001) which
has much higher strength and stiffness than the metal foams
(Deshpande et al., 2001). In the real situation, the joints are
relatively weak and easily broken under loading. After joint
Fig. 13 – Relationship between the forming pressure and
breaking the sliding between cross wires occurs under the
the relative density for the entangled Ti wire materials.
elevated loading. The sliding mechanism raises an important
factor in this material: internal friction, which causes some
novel mechanical behaviors, such as elastic hysteresis (Liu
et al., 2010; Tan et al., 2009) and high capacity of energy
absorption (Tan et al., 2009).
Considering the entangled wire materials without sinter-
ing (joint does not form), the internal friction force for one
cross point can be expressed as:
f
Fi = µFni (8)
f
where Fi denotes the friction force of a cross point. µ is the
coefficient of friction. Fni denotes normal force exerted by
each wire on the other, which can be related to the forming
pressure (denoted as PF ) of the entangled titanium wire
materials as listed in Table 1. Since the forming pressure is
balanced by the total internal contacted stress during forming
processing, it is reasonable to assume:
Fig. 14 – Relationship between the yield strength and the
nFni ≈ PF (9) relative density for the entangled Ti wire materials.
where n is the numbers of the cross points per unit
area perpendicular to the stress direction. If the sliding
The strength predicted by this simple equation is in good
mechanism dominates (wire deformation resistance can be
agreement with the measured yield strength of the entangled
ignored in this case), the deformation resistance (equals to
titanium wire materials as shown in Fig. 14, indicating that
the strength) of the entangled wire materials can be simply
sliding mechanism and internal friction dominate during
taken as the total internal friction force:
deformation in this material.
f
σ ≈ nFi = nµFni . (10) In order to evaluate the mechanical properties of the
porous materials, Deshpande et al. (2001) draw Hashin–
By plotting the relative density against the forming
Shtrikman upper bounds (Hashin and Shtrikman, 1963) on
pressure based on the measurements (columns 2 and 3 in
the stiffness and the uniaxial yield strength obtained by
Table 1) and Eq. (3), an empirical relationship between the two
transforming the linear HS bound to the perfectly plastic
values can be obtained (as shown in Fig. 13):
case using the prescription of Suquet (1993). They found
ρ
 ∗
that the stiffness and strength of the octet-truss lattice
ln PF = 1.366 + 7.365 . (11)
ρs material were about half the theoretical upper bound values
Thus, a relationship between the strength and the relative for relative density between 0.01 and 0.1, and exceeded the
density can be established: corresponding values for metallic foams by a factor between 3
and 10. It is convenient to compare the mechanical properties
ρ
 ∗
ln σ = C + 7.365 (12) of the entangled titanium wire materials with those of other
ρs
porous materials by using the theoretical HS upper bounds as
where C is a constant determined by the coefficient of friction shown in Fig. 15. The strength of the entangled titanium wire
of titanium wires: C = ln µ + 1.366. For dry friction of titanium materials is about half the theoretical upper bound values for
wires, µ ≈ 0.2–0.3, then, C ≈ 0. Thus, relative density between 0.3 and 0.6, which are comparable to
ρ
 ∗ that of the octet-truss lattice material. But in lower relative
ln σ = 7.365 . (13) density (below 0.3), the strength is far below the theoretical
ρs
28 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

Fig. 15 – Comparison among the stiffness and strength of


the entangled Ti wire materials, other porous materials
and the upper bounds for isotropic voided materials.

upper bound values. Even so, the relative strength of the


entangled titanium wire materials is higher than those of
metal foams if the relative density is not considered. Typically,
the relative elastic modulus of the entangled titanium wire
materials is far below the theoretical upper bound values, and
is also far below that of the octet-truss lattice material. This
is because of the very flexible entangled wire structure.
Fig. 16 – Effect of the forming pressure on the wire cross
4.2. Strengthening of the entangled titanium wire joints. (a) with 81.5 MPa, and (b) with 180.4 MPa.
materials

Unlike other metallic porous materials, the entangled a relative strong joint is expected. These observations help us
titanium wire materials behave with sliding and friction to understand why the strength and stiffness of the entangled
mechanism under loading. In order to strengthen the titanium wire materials increase significantly as the density
materials one must block the wire sliding. Sintering leads increases (Fig. 15).
to form joints at the wire cross points which prevent wire It is undoubted that the spatial architecture of the en-
sliding, thus improve stiffness and strength. Fig. 12(c) has tangled titanium wire materials plays an important role in
visually illustrated this efficiency of the sintering. the mechanical behavior and properties. We compared the
The present work has confirmed that sintering strengthen- non-woven structure with a quasi-ordered entangled wire
ing is effective for the entangled wire materials. All the sin- structure, and found distinct difference in the mechanical
tering parameters (e.g., temperature, holding time, protective properties (Liu et al., 2010; Tan et al., 2009). The quasi-
argon atmosphere, and following heat treatment) are set for
ordered entangled wire materials are more flexible and well-
forming strong joints. Even so, most of the joints are weak
proportioned in the porous structure. These are favorable for
in practical situation compared to the wire itself. They will
the biomedical applications. But their stiffness is still low.
preferentially fail under loading. It is desired that most the
Strengthening is highly needed when they are used for ar-
joints are as strong as possible, so that the entangled wire
tificial hip joint (Garrett et al., 2006; Long and Rack, 1998) or
materials are strong enough for applications. How to make
spine fusion (Imwinkelried, 2007; Müller et al., 2006; Singh
stronger joint is a problem in practice. The strength of the
et al., 2010) where the stiffness must be comparable to those
sintered joints depends on the pre-pressure (local contacted
of the cortical bone. Forming strong wire cross joints in the
stress), contacted area, and the sintering parameters. Both
entangled wire structure should be an effective strengthen-
the pre-pressure and the contacted area are determined by
the forming pressure as listed in Table 1, which can be re- ing method.
lated to the porosity of the material. The high forming pres- For the non-woven entangled titanium wire materials, the
sure corresponds to the high pre-pressure and large contacted microstructural homogeneity is another concern. The hetero-
area at the cross point. Fig. 16 shows cross points formed by geneous entangled wire structure leads to scattered mechan-
using two different forming pressures. It is evident that form- ical properties and undesirable pore distribution. Although
ing pressure has caused local plastic deformation at the cross a great effort has been made in making homogeneous non-
points. A forming pressure of 81.5 MPa leads to small con- woven entangled wire materials, some abnormal large voids
tacted area (Fig. 16(a)), thus a relative weak joint is expected; or local textures still exist in the materials (see section views
but a forming pressure of 180.4 MPa results in severe local as shown in Fig. 2). Alternatively, the quasi-ordered entan-
plastic deformation and large contacted area (Fig. 16(b)), thus gled wire materials are practically useful because they are
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 29

increase as the porosity decreases. The elastic modulus is


very low for this kind of material because of the structural
flexibility. Typically, 44.7% porosity corresponds to 75 MPa
yield strength, 108 MPa ultimate tensile strength, and 1.05 GPa
elastic modulus; while 57.9% porosity corresponds to only
24 MPa yield strength, 47.5 MPa ultimate tensile strength, and
0.33 GPa elastic modulus.
The entangled titanium wire materials are very promis-
ing for implant applications because of their very good
toughness, perfect flexibility, high strength, adequate elastic
modulus, and low cost. But some modification and improve-
ment in the average pore size and/or the balanced mechanical
Fig. 17 – Entangled metallic wire materials with improved
properties are needed when used for different applications.
homogeneity.
For example, when they are used for bone ingrowth, a rel-
ative larger diameter titanium wire is recommended which
more homogeneous in porous structure and easily mass- results in larger average pore size. Alternatively, the larger
manufactured (Tan et al., 2009). Fig. 17 shows some cylindrical porosity also leads to larger average pore size, but results in
samples of the quasi-ordered entangled wire materials with low strength and small elastic modulus. Generally, the elastic
different porosities. We will endeavor to strengthen them modulus of the entangled titanium wire materials is compa-
with various methods in future study, so as to meet the de- rable to that of the cancellous bone, but lower than the corti-
mand of the load-bearing biomedical applications. cal bone. It is possible to enhance the stiffness of this material
via the entangled wire structure design and the strength-
ening of the wire cross joints. All these need comprehen-
5. Conclusions and remarks sive research on the structural mechanics and the fabrication
procedures of this kind of material.
The entangled titanium wire materials with non-woven
structure can be made under forming pressure of 12–180 MPa.
The porosity of the entangled titanium wire materials is Acknowledgments
determined by the forming pressure. It decreases as the
forming pressure increases. Typically, the 12.6 MPa forming We wish to thank Prof. John Banhart and Dr. García
pressure corresponds to 82% porosity, while the 180.4 MPa Moreno, from Hahn–Meitner-Institut Berlin, Germany, for
corresponds to 48% porosity. their assistance in the tomography experiments and helpful
The pores in the entangled titanium wire materials discussions.
are irregular in shape, which have a nearly half-normal
distribution in size range. The average pore size depends on REFERENCES
the porosity and the wire diameter. The larger porosity and
the larger wire diameter lead to the larger average pore size.
More than 50% pores are in the range of 50–200 µm for all the Albrektsson, T., Johansson, C., 2001. Osteoinduction, osteocon-
materials tested in this study, but only when the porosity is duction and osseointegration. Eur. Spine J. 10, S96–S101.
larger than 60%, near 50% pores are larger than 100 µm for the Banhart, J., 2001. Manufacture, characterisation and application of
cellular metals and metal foams. Prog. Mater. Sci. 46, 559–632.
sample made of 0.08 mm-diameter wire. For example, the 82%
Bréme, J., Biehl, V., Schulte, W., d’Hoedf, B., Donath, K., 1993.
porosity corresponds to about 75% pores above 100 µm and Development and functionality of isoelastic dental implants
25% pores under 100 µm; while the 48% porosity corresponds of titanium alloys. Biomaterials 14, 887–892.
to about 70% pores under 100 µm. Bungo, O., Mitsuru, T., Shunsuke, F., Masashi, N., Tadashi, K.,
The entangled titanium wire materials exhibit initial near- Takashi, N., 2006. Pore throat size and connectivity determine
linear elastic stage, nonlinear plastic deformation stage, and bone and tissue ingrowth into porous implants: three-
failure stage under uniaxial tensile loading. The near-linear dimensional micro-CT based structural analyses of porous
bioactive titanium implants. Biomaterials 27, 5892–5900.
elastic stage corresponds to the structure elastic deformation;
Cachinho, S.C., Correia, R.N., 2008. Titanium scaffolds for
the nonlinear plastic deformation stage corresponds to the
osteointegration: mechanical, in vitro and corrosion behavior.
structure plastic deformation via wires sliding over each J. Mater. Sci. Mater. Electron. 19, 451–457.
other. When the material is in small porosity the sliding Chang, Y.S., Oka, M., Kobayashi, M., Gu, H.O., Li, Z.L., Nakamura,
can be blocked, so that the structure plastic deformation T., Ikada, Y., 1996. Significance of interstitial bone ingrowth
becomes unnoticeable. The failure mode of the materials can under load-bearing conditions: a comparison between solid
be entangled wire loosening or wires broken, which depends and porous implant materials. Biomaterials 17, 1141–1148.
on the porosity. The material with large porosity fails via wire Deshpande, V.S., Fleck, N.A., Ashby, M.F., 2001. Effective properties
of the octet-truss lattice material. J. Mech. Phys. Solids 49,
loosening; otherwise it fails via wires broken.
1747–1769.
The entangled titanium wire materials exhibit distinct Dolder, J., Farber, E., Spauwen, P.H.M., Jansen, J.A., 2003. Bone
yielding that corresponds to startup of the wire sliding tissue reconstruction using titanium fiber mesh combined
in large scale. The yield strength significantly depends on with rat bone marrow stromal cells. Biomaterials 24,
the porosity. Both the yield strength and ultimate strength 1745–1750.
30 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31

Donald, T.R., Albert, H.B., Cleveland, O., 1974. The mechanical Lakes, R., 1993. Advances in negative Poisson’s ratio materials.
properties of cortical bone. J. Bone Joint Surg. Am. 56, Adv. Mater. 5, 293–296.
1001–1022. Lee, J.H., Kim, H.E., Koh, Y.H., 2009. Highly porous titanium
Ducheyne, P., Martens, M., Aernoudt, E., Mulier, J., Meester, P., (Ti) scaffolds with bioactive microporous hydroxyapatite/TiO2
1974. Skeletal fixation by metal fiber coating of the implant. hybrid coating layer. Mater. Lett. 63, 1995–1998.
Acta Orthop. Belg. 40, 799–805. Li, J.P., Li, S.H., Van Blitterswijk, C.A., De Groot, K., 2005. A
Eigler, N.L., Khorsandi, M.J., Forrester, J.S., Fishbein, M.C., Litvack, novel porous Ti6 Al4 V: characterization and cell attachment.
F., 1993. Implantation and recovery of temporary metallic J. Biomed. Mater. Res. Part A 73, 223–233.
stents in canine coronary arteries. J. Am. Coll. Cardiol. 22, Li, J.P., Pamela, H., Mirella, D., Clayton, E.W., Joost, R.W., Clemens,
1207–1213. A.B., Klaas, G., 2007. Bone ingrowth in porous titanium
Freyman, T.M., Yannas, I.V., Gibson, L.J., 2001. Cellular materials implants produced by 3D fiber deposition. Biomaterials 28,
as porous scaffolds for tissue engineering. Prog. Mater. Sci. 46, 2810–2820.
273–282. Li, J.P., Wijn, J.R., Blitterswijk, C.A., Groot, K., 2006. Porous Ti6 Al4 V
Fujibayashi, S., Nakamura, T., Nishiguchi, S., Tamura, J., Uchida, scaffold directly fabricating by rapid prototyping: preparation
M., Kim, H.M., Kokubo, T., 2001. Bioactive titanium: effect and in vitro experiment. Biomaterials 27, 1223–1235.
of sodium removal on the bone-bonding ability of bioactive Liu, P., 2010. Fabrication and mechanical behavior of the high-
titanium prepared by alkali and heat treatment. J. Biomed. tough entangled metal wire materials and their application for
Mater. Res. 56, 562–570. implant. Doctoral Dissertation. Shanghai Jiaotong University,
Fujibayashi, S., Neo, M., Kim, H.M., Kokubo, T., Nakamura, T., Shanghai.
2004. Osteoinduction of porous bioactive titanium metal. Liu, P., He, G., Wu, L.H., 2009. Uniaxial tensile stress–strain
Biomaterials 25, 443–450. behavior of entangled steel wire material. Mater. Sci. Eng. A
Garcia, M.F., Kardjilov, N., 2005. Cellular architecture—overview of 509, 69–75.
X-ray and neutron radioscopy and tomography. Mater. World Liu, P., Tan, Q.B., Wu, L.H., He, G., 2010. Compressive and
6. pseudo-elastic hysteresis behavior of entangled titanium wire
Garrett, R., Abhay, P., Panagiotis, A.D., 2006. Fabrication methods materials. Mater. Sci. Eng. A 527, 3301–3309.
Long, M., Rack, H.J., 1998. Titanium alloys in total joint
of porous metals for use in orthopaedic applications.
replacement—a materials science perspective. Biomaterials
Biomaterials 27, 2651–2670.
19, 1621–1639.
Geetha, M., Singh, A.K., Asokamani, R., Gogia, A.K., 2009. Ti based
Manukyana, K., Amirkhanyan, N., Aydinyan, S., Danghyan, V.,
biomaterials, the ultimate choice for orthopaedic implants—a
Grigoryan, R., Sarkisyan, N., Gasparyan, G., Aroutiounian, R.,
review. Prog. Mater. Sci. 54, 397–425.
Kharatyan, S., 2010. Novel NiZr-based porous biomaterials:
Gibson, L.J., Ashby, M.F., 1997. Cellular Solids: Structure and
Synthesis and in vitro testing. Chem. Eng. J. 162, 406–414.
Properties, 2nd ed. Cambridge University Press, Cambridge,
Mullen, L., Stamp, R.C., Brooks, W.K., Jones, E., Sutcliffe, C.J., 2008.
UK.
Selective laser melting: a regular unit cell approach for the
Hashin, Z., Shtrikman, S., 1963. A variational approach to the
manufacture of porous, titanium, bone ingrowth constructs,
theory of the elastic behaviour of multi-phase materials. J.
suitable for orthopedic applications. J. Biomed. Mater. Res. B
Mech. Phys. Solids 11, 127–140.
Appl. Biomater. 89B, 325–334.
Hollander, D.A., Walter, M., Wirtz, T., Sellei, R., Schmidt-Rohlfing,
Müller, U., Imwinkelried, T., Horst, M., Sievers, M., Graf-Hausner,
B., Paar, O., Erli, H.J., 2006. Structural, mechanical and in
U., 2006. Do human osteoblasts grow into open-porous
vitro characterization of individually structured Ti–6Al–4V
titanium? Eur. Cells Mater. 11, 8–15.
produced by direct laser forming. Biomaterials 27, 955–963. Niinomi, M., 2008. Mechanical biocompatibilities of titanium
Hollister, S.J., 2005. Porous scaffold design for tissue engineering. alloys for biomedical applications. J. Mech. Behav. Biomed.
Nature Mater. 4, 518–524. Mater. 1, 30–42.
Holy, C.E., Shoichet, M.S., Davies, J.E., 2000. Engineering three- Nishiguchi, S., Nakamura, T., Kobayashi, M., Kim, H.M., Miyaji, F.,
dimensional bone tissue in vitro using biodegradable scaffolds: Kokubo, T., 1999. The effect of heat treatment on bone-bonding
investigating initial cell-seeding density and culture period. ability of alkali-treated titanium. Biomaterials 20, 491–500.
J. Biomed. Mater. Res. 51, 376–382. Oh, I.H., Nomura, N., Masahashi, N., Hanada, S., 2003. Mechanical
Imwinkelried, T., 2007. Mechanical properties of open-pore properties of porous titanium compacts prepared by powder
titanium foam. J. Biomed. Mater. Res. Part A 81, 964–970. sintering. Scr. Mater. 49, 1197–1202.
Jacobs, J.J., Sumner, D.R., Galante, J.O., 1993. Mechanisms of bone Paquay, Y.C., Ruijter, J.E., Waerden, J.P., Jansen, J.A., 1997. Wound
loss associated with total hip replacement (Review). Orthop. healing phenomena in titanium fibre mesh: the influence of
Clin. North Am. 24, 583–590. the length of implantation. Biomaterials 18, 161–166.
Kashef, S., Asgari, A., Hilditch, T.B., Yan, W., Goel, V.K., Hodgson, Singh, R., Lee, P.D., Lindley, T.C., Kohlhauser, C., Hellmich, C.,
P.D., 2010. Fracture toughness of titanium foams for medical Bram, M., Imwinkelried, T., Dashwood, R.J., 2010. Character-
applications. Mater. Sci. Eng. A 527, 7689–7693. ization of the deformation behavior of intermediate poros-
Kim, H.M., Kokubo, T., Fujibayashi, S., Nishiguchi, S., Nakamura, ity interconnected Ti foams using micro-computed tomog-
T., 2000. Bioactive macroporous titanium surface layer on raphy and direct finite element modeling. Acta Biomater. 6,
titanium substrate. J. Biomed. Mater. Res. 52, 553–557. 2342–2351.
Körner, C., Singer, R., 2000. Processing of metal foams—challenges Stevens, M.M., 2008. Biomaterials for bone tissue engineering.
and opportunities. Adv. Eng. Mater. 2, 159–165. Mater. Today 11, 18–25.
Krishna, B.V., Bose, S., Bandyopadhyay, A., 2007. Low stiffness Suquet, P.M., 1993. Overall potentials and extremal surfaces of
porous Ti structures for load-bearing implants. Acta Biomater. power law or ideally plastic composites. J. Mech. Phys. Solids
3, 997–1006. 41, 981–1002.
Krishna, B.V., Xue, W., Bose, S., Bandyopadhyay, A., 2008. Tan, Q.B., Liu, P., Du, C.L., Wu, L.H., He, G., 2009. Mechanical
Engineered porous metals for implants. JOM 60, 45–48. behaviors of quasi-ordered entangled aluminum alloy wire
Laheurte, P., Prima, F., Eberhardt, A., Gloriant, T., Wary, M., Patoor, material. Mater. Sci. Eng. A 527, 38–44.
E., 2010. Mechanical properties of low modulus beta titanium Taylor, R.P., McClain, S.T., Berry, J.T., 1999. Uncertainty analysis
alloys designed from the electronic approach. J. Mech. Behav. of metal-casting porosity measurements using Archimedes’
Biomed. Mater. 3, 565–573. principle. Int. J. Cast Met. Res. 11, 247–257.
J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M AT E R I A L S 5 (2012) 16–31 31

Traini, T., Mangano, C., Sammons, R.L., Mangano, F., Macchi, Wallach, J.C., Gibson, L.J., 2001. Mechanical behavior of a three-
A., Piattelli, A., 2008. Direct laser metal sintering as a new dimensional truss material. Internat. J. Solids Struct. 38,
approach to fabrication of an isoelastic functionally graded 7181–7196.
material for manufacture of porous titanium dental implants. Warren, W.E., Kraynik, A.M., 1991. The nonlinear elastic behavior
Dent. Mater. 24, 1525–1533. of open-cell foams. J. Appl. Mech. 58, 376–381.
Vehof, J.W.M., Haus, M.T.U., Ruijter, A.E., Spauwen, P.H.M., Jansen, Wen, C.E., Mabuchi, M., Yamada, Y., Shimojima, K., Chino, Y.,
J.A., 2002. Bone formation in transforming growth factor beta- Asahina, T., 2001. Processing of biocompatible porous Ti and
1-loaded titanium fiber mesh implants. Clin. Oral. Implan. Res. Mg. Scr. Mater. 45, 1147–1153.
13, 94–102. Zhao, J., Lu, X., Weng, J., 2008. Macroporous Ti-based
Vehof, J.W.M., Spauwen, P.H.M., Jansen, J.A., 2000. Bone formation composite scaffold prepared by polymer impregnating
in calcium-phosphate-coated titanium mesh. Biomaterials 21, method with calcium phosphate coatings. Mater. Lett. 62,
2003–2009. 2921–2924.

Das könnte Ihnen auch gefallen