mechanical behavior and microstructure of compressed ti foams …heeman/paper/mechanical... ·...

10
www.elsevier.com/locate/jmbbm Available online at www.sciencedirect.com Research Paper Mechanical behavior and microstructure of compressed Ti foams synthesized via freeze casting Pe ´ter Jenei a , Hyelim Choi b , Adria ´n To ´th a , Heeman Choe b , Jeno ˝ Gubicza a,n a Department of Materials Physics, Eötvös Loránd University, P.O.B. 32, Budapest H-1518, Hungary b School of Advanced Materials Engineering, Kookmin University, 77 Jeongneung-ro, Seongbuk-gu, Seoul 136-702, Republic of Korea article info Article history: Received 6 April 2016 Received in revised form 13 June 2016 Accepted 9 July 2016 Available online 15 July 2016 Keywords: Ti-foam Freeze-casting Plastic deformation Young's modulus Yield strength Dislocations abstract Pure Ti and Ti5%W foams were prepared via freeze casting. The porosity and grain size of both the materials were 3233% and 1517 mm, respectively. The mechanical behavior of the foams was investigated by uniaxial compression up to a plastic strain of 0.26. The Young's moduli of both foams were 23 GPa, which was in good agreement with the value expected from their porosity. The Young's moduli of the foams were similar to the elastic modulus of cortical bones, thereby eliminating the osteoporosis-causing stress-shielding effect. The addition of W increased the yield strength from 196 MPa to 235 MPa. The microstructure evolution in the grains during compression was studied using electron backscatter diffrac- tion (EBSD) and X-ray line prole analysis (XLPA). After compression up to a plastic strain of 0.26, the average dislocation densities increased to 3.4 10 14 m 2 and 5.9 10 14 m 2 in the Ti and TiW foams, respectively. The higher dislocation density in the TiW foam can be attributed to the pinning effect of the solute tungsten atoms on dislocations. The experimentally measured yield strength was in good agreement with the strength calculated from the dislocation density and porosity. This study demonstrated that the addition of W to Ti foam is benecial for biomedical applications, because the compressive yield strength increased while its Young's modulus remained similar to that of cortical bones. & 2016 Elsevier Ltd. All rights reserved. 1. Introduction Metallic materials have signicant advantages over ceramic and polymer materials as orthopedic implants owing to their excellent properties such as superior strength, fracture toughness and ductility (Geetha et al., 2009; Ryan et al., 2006). Particularly, signicant attention has been paid to Ti and Ti-based alloys owing to their relatively low modulus, http://dx.doi.org/10.1016/j.jmbbm.2016.07.012 1751-6161/& 2016 Elsevier Ltd. All rights reserved. Abbreviations: EBSD, electron backscatter diffraction; XLPA, X-ray line prole analysis; SEM, scanning electron microscope; LAGB, low-angle grain boundary n Corresponding author. Fax: þ36 1 372 2811. E-mail address: [email protected] (J. Gubicza). journal of the mechanical behavior of biomedical materials 63(2016)407–416

Upload: others

Post on 31-Jan-2020

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6

http://dx.doi.org/101751-6161/& 2016 El

Abbreviations: EB

LAGB, low-angle gnCorresponding autE-mail address:

Research Paper

Mechanical behavior and microstructureof compressed Ti foams synthesizedvia freeze casting

Peter Jeneia, Hyelim Choib, Adrian Totha, Heeman Choeb, Jeno Gubiczaa,n

aDepartment of Materials Physics, Eötvös Loránd University, P.O.B. 32, Budapest H-1518, HungarybSchool of Advanced Materials Engineering, Kookmin University, 77 Jeongneung-ro, Seongbuk-gu, Seoul 136-702,Republic of Korea

a r t i c l e i n f o

Article history:

Received 6 April 2016

Received in revised form

13 June 2016

Accepted 9 July 2016

Available online 15 July 2016

Keywords:

Ti-foam

Freeze-casting

Plastic deformation

Young's modulus

Yield strength

Dislocations

.1016/j.jmbbm.2016.07.012sevier Ltd. All rights rese

SD, electron backscatt

rain boundaryhor. Fax: þ36 1 372 [email protected]

a b s t r a c t

Pure Ti and Ti–5%W foams were prepared via freeze casting. The porosity and grain size of

both the materials were 32–33% and 15–17 mm, respectively. The mechanical behavior of the

foams was investigated by uniaxial compression up to a plastic strain of �0.26. The Young's

moduli of both foams were �23 GPa, which was in good agreement with the value expected

from their porosity. The Young's moduli of the foams were similar to the elastic modulus of

cortical bones, thereby eliminating the osteoporosis-causing stress-shielding effect. The

addition of W increased the yield strength from �196 MPa to �235 MPa. The microstructure

evolution in the grains during compression was studied using electron backscatter diffrac-

tion (EBSD) and X-ray line profile analysis (XLPA). After compression up to a plastic strain of

�0.26, the average dislocation densities increased to �3.4� 1014 m�2 and �5.9� 1014 m�2 in

the Ti and Ti–W foams, respectively. The higher dislocation density in the Ti–W foam can be

attributed to the pinning effect of the solute tungsten atoms on dislocations. The

experimentally measured yield strength was in good agreement with the strength calculated

from the dislocation density and porosity. This study demonstrated that the addition of W to

Ti foam is beneficial for biomedical applications, because the compressive yield strength

increased while its Young's modulus remained similar to that of cortical bones.

& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Metallic materials have significant advantages over ceramic

and polymer materials as orthopedic implants owing to their

rved.

er diffraction; XLPA, X-

(J. Gubicza).

excellent properties such as superior strength, fracture

toughness and ductility (Geetha et al., 2009; Ryan et al.,

2006). Particularly, significant attention has been paid to Ti

and Ti-based alloys owing to their relatively low modulus,

ray line profile analysis; SEM, scanning electron microscope;

Page 2: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6408

high strength, superior corrosion resistance, and excellentbiocompatibility (through the formation of an oxide layerupon contact with air) (Yamamoto et al., 2012; Long and Rack,1998; Choi et al., 2014).

Despite the increasing reputation of Ti and Ti-based alloysas orthopedic implants, they still suffer from stress-shieldingeffect when used in the bulk form because they havesignificantly greater elastic moduli than that of bone, thuseventually resulting in osteoporosis and loosening of theimplant (elastic modulus for natural bone: 3–20 GPa; elasticmodulus for Ti and Ti-based alloys: 55–117 GPa) (Geetha et al.,2009; Li et al., 2004; Krishna et al., 2007). Therefore, theirporous counterparts with lower elastic moduli values areoften preferred for orthopedic implant applications. More-over, the porous structure of the implant plays an importantrole in bone integration, i.e., the porous surface facilitatesstrong interlocking with the bone tissue around the implant,resulting in high resistance to fatigue loading and biomecha-nical compatibility (Geetha et al., 2009; Sauer et al., 1974;Niinomi, 2008).

Several manufacturing methods have already been pro-posed for porous Ti. One of the most common methods is thespace-holder technique. In the space-holder method, Tipowder is mixed with organic solvents and carbamide asspace-holder, which is removed later by heat-treatment toleave hollow spaces (Niu et al., 2009). A gel-casting is some-what similar to the space-holder method. A mixture of Tipowder and some solvents forms a porous structure throughcasting and gelation, subsequently followed by drying andsintering (Erk et al., 2008). Another interesting process is theprinting processing in which Ti powder and solvent aremixed to produce Ti ink and fabricate a 3-dimensional porousstructure of Ti foam through sintering (Hong et al., 2011).Finally, Ibrahim et al. also demonstrated the processing ofporous Ti using spark plasma sintering (Ibrahim et al., 2011).In this study, we selected a freeze-casting method to produceTi and Ti–5%W alloy foams, because this processing methodcan control the pore morphology, pore size, and porosity of Tifoams fairly reasonably. Therefore, the freeze-casting methodcan allow relatively easy scale-up and commercialization.

Along with the elastic modulus, we must also investigatethe plastic properties and fundamental deformation mechan-isms for the porous implant materials because those proper-ties are critical for their successful use as load-bearingimplant. For example, the femur bone is normally expectedto support approximately 30 times the weight of a typicaladult female body (Magee, 2008; D’Angeli et al., 2013). How-ever, only a few studies have focused on comprehensiveanalysis of mechanical properties of porous Ti and Ti-basedalloys from the perspective of their potential use in biome-dical applications. For instance, the hardness, compressivestrength, and stiffness (Young's modulus) were analyzedthrough compressive tests on porous Ti and Ti-based alloys(Taniguchi et al., 2016; Hong et al., 2011; Muñoz et al., 2015).Additionally, a fatigue test was performed on strain accumu-lated Ti–6Al–4V foam with the corresponding images andmodeling results analyzed on the tested samples (Zhao et al.,2016). Despite these studies on compressive strength, fatigue,and fracture, a systematic analysis is still required particu-larly on the variations in microstructural and physical

properties such as variations in pore morphology, deforma-

tion mechanism, and elastic modulus during compression of

porous Ti and Ti-based alloys.Therefore, in the present study, we synthesized Ti and Ti–

5%W alloy foams via freeze casting for a systematic analysis

on the microstructural evolution of the Ti foams during

compression. Tungsten (W) was added to improve the

strength and wear resistance through a solid-solution

strengthening effect of W in Ti grains (Choi et al., 2014,

Frary et al., 2003). Moreover, we analyzed the microstructural

evolution, elastic moduli, compressive strengths and defor-

mation mechanisms of the freeze-cast porous Ti and Ti–5%W

alloy foams using compression test, electron backscatter

diffraction (EBSD) and X-ray line profile analysis (XLPA). In

particular, we investigated the subgrain boundaries and the

dislocation densities of the compressed samples using EBSD

and XLPA. We also examined the effect of porosity and

dislocations on the mechanical properties (e.g., elastic mod-

ulus, flow stress) of Ti foams. This systematic study using a

range of analytical test methods are expected to provide

valuable insights on the mechanical and deformation beha-

vior of porous Ti and Ti-based alloy foams and other applic-

able porous implants under complex stress and strain states

for their potential use as biomedical materials.

2. Material and methods

2.1. Preparation of Ti foams via freeze casting

Pure Ti foam was synthesized from commercially pure Ti

powder (Alfa Aesar, MA, USA) with a mesh value of 325

(particle size is smaller than 44 mm). The concentration of

metallic impurities was less than 0.2%. Among the nonme-

tallic elements, oxygen and nitrogen have the highest con-

centrations with values of 0.694% and 0.3%, respectively,

according to the manufacturer's analysis. Ti–W alloy foam

was also synthesized. In this case, 5 wt% W powder with an

average particle size of 1 mm was added to the Ti powder. A

sequence of experimental procedures was performed to

prepare the samples prior to freeze casting. First, polyvinyl

alcohol (PVA; Mw 89,000–98,000, purity �99%, Sigma-Aldrich

Co., MO, USA) was dissolved in distilled water, and an

appropriate amount of Ti or Ti/W powder mixture was then

added to the prepared solution to complete the slurry. The

slurry was then poured directly on top of a Cu chiller rod of

40 mm in diameter standing in a stainless steel vessel under

liquid N2. In order to ensure thermal protection along the

horizontal direction and allow heat transfer through contact

between the slurry and the Cu rod, the slurry was wrapped

with polystyrene foil and inserted into a polymer mold. The

freezing rate was �10 1C/min. The frozen green body was

lyophilized in a freeze dryer (Operon, OPR-FDU-7003, Republic

of Korea) to remove ice through sublimation at �90 1C and

5�10�3 Torr for over 24 h. The lyophilized green body was

then sintered in a vacuum furnace at 1050 1C for 6 h.

Page 3: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6 409

2.2. Mechanical test

Cuboid specimens with dimensions of 3�3�6 mm3 were cutfrom the pure Ti and Ti–W foams. Quasi-static uniaxialcompression tests were conducted on these samples at astrain rate of �10�4 s�1 using an MTS universal mechanicaltesting machine. Three specimens were tested for eachcompression test and the results showed good reproducibil-ity. The mechanical tests were interrupted at several selectedstrains in order to investigate the evolution of the defectstructure during compression. The true strain was deter-mined from the cross-head displacement as ln(l0/l), where l0and l are the heights of the initial and deformed samples,respectively. The true stress was calculated as the ratio of theforce and the cross-sectional area of the compressed sample.For porous structures, the actual cross-sectional area cannotbe determined from the cross-section of the initial sampleand the true strain, because the sample volume may changeduring compression. Therefore, the cross-sectional area wasmeasured directly at the beginning and end of deformation aswell as at the strains where the compression test wasinterrupted. Between these strain values, the cross-sectionwas determined by interpolation using an exponentialfunction.

2.3. Microstructure analysis by EBSD and XLPA

The EBSD analysis was performed using an FEI Quanta 3Ddual beam scanning electron microscope (SEM) equippedwith an EDAX type EBSD system. In order to obtain high-quality EBSD images, mechanically polished surfaces of thesamples were ion milled for 25 min by a 10 keV Ar-ion beamat an incidence angle of �61. The EBSD scans covered200 μm�200 μm regions with a step size of 0.5 μm betweenneighboring measurement positions. The grain size andmisorientation distributions were determined from the EBSDscans using OIM software (version 5.3) from TexSEMLaboratories (TSL).

Moreover, XLPA was also performed to investigate themicrostructure of the samples. The X-ray line profiles weremeasured with a high-resolution rotating anode diffract-ometer (Rigaku, RA MultiMax-9) using CuKα1 (λ¼0.15406 nm)radiation. Two-dimensional (2D) imaging plates were used fordetecting the Debye–Scherrer diffraction rings. The line pro-files were obtained by integrating the 2D intensity distribu-tion along the rings, and the X-ray diffraction patterns wereevaluated using the convolutional multiple whole profile(CMWP) fitting method (Ribárik et al., 2004). In this procedure,the diffraction pattern is fitted by the sum of a backgroundspline and the convolution of the instrumental pattern andthe theoretical line profiles related to the crystallite size anddislocations. The instrumental profiles were measured on aLaB6 standard sample under the same conditions as thoseapplied for the Ti foam samples. The theoretical profilefunctions were calculated on the basis of a model of themicrostructure having spherical crystallites and lognormalsize distribution, and the lattice strains were assumed to becaused by dislocations. Approximately fourteen peaks of Tiwere fitted in the evaluation of line profiles. The area-weighted mean crystallite size (ox4area) and the dislocation

density (ρ) were obtained from this method. In addition, theprevailing dislocation slip systems in hexagonal crystals weredetermined. The experimental values of parameters q1 and q2in the dislocation contrast factors obtained by the CMWPmethod depend on the type of dislocations. The theoretical q1and q2 values for the 11 possible slip systems in Ti werecalculated according to Kužel and Klimanek (1988) and arelisted in Dragomir and Ungár (2002). The 11 dislocation slipsystems can be classified into 3 groups on the basis of theirBurgers vectors: b1 ¼ 1=3⟨2110⟩ (oa4 type), b2 ¼ ⟨0001⟩ (oc4type), and b3 ¼ 1=3⟨2113⟩ (ocþa4 type). A comparison of theexperimental and theoretical values of parameters q1 and q2yields the fractions of oa4, oc4, and ocþa4 dislocations. Adetailed description of the evaluation procedure can be foundin Máthis et al. (2004).

3. Results

3.1. Microstructures of initial foams

The initial microstructures of the pure Ti and Ti–W foams areshown in Figs. 1a and 2a, respectively. The black areascorrespond to the pores, while the colored areas indicatethe orientations of the grains. The volume fraction of poreswas determined as the fraction of black areas in the SEMimages. According to quantitative metallography, these twofractions are equal within error range. The pore volumefraction values for the pure Ti and Ti–W foams were 33%and 32%, respectively, thus indicating that the porosity wasnot altered by W alloying. The porosity was also estimatedfrom the density obtained as the ratio of mass to volume ofthe samples. A direct Archimedes' method could not beemployed for determining the volume of these foamsbecause they have a large fraction of open pores, as revealedby microscopic observations. Alternatively, the volume wascalculated from the product of the three orthogonal dimen-sions of the cuboid specimens. The porosity volume fractionsfor Ti and Ti–W foams obtained from the density measure-ment were 33% and 30%, respectively, which are in goodagreement with the values obtained from EBSD analysis. Theaverage grain size values for the Ti and Ti–W foams were17 mm and 15 mm, respectively, therefore neither the grainsize nor the porosity was significantly altered in the twofoams. Electron probe microanalysis (not presented in thispaper) indicated that in the Ti–W foam, �2 wt% W wascompletely dissolved into Ti grains, while the other 3 wt%W remained uniformly as secondary phase particles.

3.2. Mechanical behavior of foams under compressive load

Fig. 3a and b show the true stress vs. true strain curves for theTi and Ti–W foams during compression. The mechanical testwas performed up to a total strain of approximately 0.3 forboth foams (which corresponds to a plastic strain of approxi-mately 0.25–0.28). Below this strain, the deformation washomogeneous in the samples because strain localization wasnot observed on the surfaces of the compressed specimens.At the end of compression tests (at the plastic strains of 0.28and 0.25 for Ti and Ti-W foams, respectively), the plastic

Page 4: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

Fig. 1 – EBSD images showing microstructural evolution in pure Ti foam (a) in as-processed state and (b), (c), and (d) afterquasi-static compression up to plastic strains of �0.06, �0.16, and �0.26, respectively. Note that the color code is shown inthe inset of (d).

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6410

Poisson's ratio was 0.3570.03 (the ratio of the magnitudes ofengineering strains measured normal and parallel to theloading axis). The mechanical tests were interrupted atplastic strains of approximately 0.06–0.07 and 0.16–0.17 toinvestigate the evolution of the defect structures (dislocationsand grain boundaries) during compression, and the resultsare presented in the following section. The Young's moduliand yield strength values of both foams were determinedfrom the first loading curve (see Table 1). It should be notedthat the compression stress–strain curves are not strictlylinear in the initial stage of deformation, and a transientstage was observed for which the derivative of the curvesincreased with increasing strain. This nonlinear stress–strainbehavior can be attributed to the nonparallel orientations ofthe contact surfaces of the samples and the instrumentcrosshead. Therefore, the Young's modulus was determinedfrom the linear section of the stress–strain curve after thistransient stage. The elastic moduli values obtained for thepure Ti and Ti–W foams were 2372 GPa and 2272 GPa,respectively, thus indicating that W alloying has negligibleinfluence on Young's modulus, which is in agreement with aprevious study where W dissolution in Ti matrix up to 10 wt%had little effect on the Young's modulus of Ti-10W alloy

(Choe et al., 2005). On the other hand, the yield strength ismore sensitive to W alloying because its value was muchhigher for the Ti–W foam (�235 MPa) than for the pure Tifoam (�196 MPa).

It should be noted that the present Ti foam processed byfreeze casting exhibited higher strength than the valuepredicted from other experiments as well as those fromtheoretical calculations (Siegkas et al., 2016). The previouslyreported experiments were carried out on Ti foams sinteredwith the help of a polymeric binder and a foaming agent. Inthe theoretical prediction, virtual foam geometries werecreated using a procedure based on Voronoi tessellation.Additionally, the quasi-static mechanical response wasdetermined by finite element simulations. More specifically,the flow stress at a compressive strain of 0.05 was predictedto be 240710 MPa from the experimental and theoreticalvalues for the foams with porosity of �33%. For our freeze-cast Ti foam, the flow stress at the same strain was measuredto be larger (�320 MPa), due most likely to the different poremorphology and structure developed in the foams caused bythe different processing method.

The strain hardening behavior of pure Ti and Ti–W foamswas analyzed by plotting the derivative of stress with respect

Page 5: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

Fig. 2 – EBSD images showing microstructural evolution in Ti–W foam (a) in as-processed state and (b), (c), and (d) after quasi-static compression up to plastic strains of �0.06, �0.16, and �0.26, respectively. Note that the color code is shown in theinset of Fig. 1d.

Fig. 3 – True stress vs. true strain plots obtained with quasi-static compression tests at room temperature for (a) pure Ti and(b) Ti–W foams.

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6 411

to plastic strain (ds/dϵp) as a function of stress (s), as shown inFig. 4 (Kocks–Mecking plot). The plastic strain was deter-mined by subtracting the elastic strain component from thetotal strain, where the former quantity was determined asthe ratio of actual stress to Young's modulus. Owing to the

interruption of compression test and the noise of measure-ment, (ds/dϵp) could not be directly obtained by determiningthe derivative of the experimental data. Therefore, thefollowing function was fitted to the s-ϵp data points and thederivative of this fitted curve with respect to plastic strain

Page 6: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

Table 1 – Porosity, Young's modulus (E), yield strength(rY), and dislocation density (ρ) at a plastic strain of �0.26for Ti and Ti–W foams, and bulk Ti.

Sample Porosity [%] E [GPa] sY [MPa] ρ [1014 m�2]

pure Ti foam 33 2272 19675 3.470.8Ti-W foam 32 2372 23576 5.970.7bulk Ti (Grade 2) 0 11473 330710 –

Fig. 4 – Derivative of stress with respect to plastic strain (dr/dϵp) as a function of stress (r) for pure Ti and Ti–W foams,and bulk Ti with Grade 2 purity level (Kocks–Mecking plot).

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6412

was then obtained as follows (Lademo et al., 1999):

s¼ s0 þ s1 1�exp �A1εp� �� �þ s2 1�exp �A2εp

� �� �

þs3 1�exp �A3εp� �� �

: ð1Þ

Fig. 3 shows that for both foams, strain hardeningdecreases gradually as stress increases, which is in accor-dance with the typical behavior of bulk Ti. The decrease in(ds/dϵp) with increasing stress is caused by the annihilationprocesses of dislocations (e.g., cross-slip and climb). Irrespec-tive of the foam state and W alloying, the (ds/dϵp) versus scurve can be divided into two segments. The decrease instrain hardening is much higher in the first segment than inthe second segment. A comparison of the two foams showsthat (ds/dϵp) is larger for Ti–W than for Ti at any stress valuebecause of the higher rate of increase in dislocation densitydue to the pinning effect of W on dislocations (see Section3.3). The significantly lower stresses and strain hardeningvalues of Ti foam than those of bulk Ti can be attributed to itsporosity.

3.3. Microstructural evolution during deformation

The microstructure evolution in the two foams was investi-gated by interrupting the compression at plastic strains ofapproximately 0.06–0.07, 0.16–0.17 and 0.25–0.28 (see Fig. 3).The latter state corresponds to the end of deformation.Although, the plastic strain values at each interrupted stateare slightly different for the two foams, we assume that thedifference is insignificant and refer these states to theapproximate plastic strain values of �0.06, �0.16, and�0.26, respectively, for both foams. Figs. 1 and 2 show the

EBSD images of the same area after compression at differentstrains for Ti and Ti–W foams, respectively. Because somegrains shifted from the study surface during compression, theevaluable area decreased with increasing strain. Therefore,the black area in the images increases regardless of thevariation in porosity. Hence, the porosity for the deformedsamples could not be evaluated from the EBSD images.Instead, the porosity values of the compressed samples wereestimated from the density obtained by measuring the massand dimensions of the specimens.

In the as-processed foams, each grain can be characterizedby a single color in the EBSD images (see Figs. 1a and 2a),indicating negligible misorientation inside the grains. On theother hand, different colors can be seen inside the grains ofthe plastically deformed samples, suggesting misorientationsbetween the different parts of the grains. It should be notedthat due to the uncertainty of the EBSD misorientationanalysis, the threshold of misorientation angle measurementwas set to be 21, a standard value in the literature. In both thecompressed foams, a majority of LAGBs (more than 95%) werecharacterized by misorientations less than 61. Fig. 5a and bshow the evolution of misorientation distributions in thepure Ti and Ti–W foams in an angle range of 21–61, respec-tively. The fraction of boundaries at a certain misorientationangle was calculated as the ratio of the length of theseboundaries to the total length of edges of pixels inside thestudied area (the possible locations of boundaries). Fig. 5a andb show that the amount of LAGBs increases with increasingplastic strain in both foams. LAGBs are typically formed bythe arrangement of dislocations into low-energy configura-tions. Therefore, the increase in dislocation density duringcompression was investigated using XLPA.

The dislocation density was determined from the breadthsof the X-ray diffraction peaks. The peak profiles wereobtained from the Debye–Scherrer diffraction rings detectedon the imaging plates. Fig. 6a and b show Debye–Scherrerrings for reflection 110 of the as-processed pure Ti-basedfoam sample and the sample compressed up to a plasticstrain of �0.26, respectively. The ring for the as-processed Tifoam consists of very narrow spots (see Fig. 6a) scatteredfrom different grains. These measured peaks are as narrow asthe instrumental broadening (Δ(2Θ)¼0.031), thus suggesting alower dislocation density and larger crystallite size whencompared with the detection limits of the present diffractionsetup (1013 m�2 for the dislocation density and 1 mm for thegrain size). In the case of the compressed sample, the numberof sharp spots decreased and a continuous broad ring wasformed (see Fig. 6b) owing to the increase in dislocationdensity and/or reduction of crystallite size. The coexistenceof both sharp spots and continuous broad ring segmentsindicates an inhomogeneous spatial distribution of latticedistortions caused by the compression of the foam. The sharpspots and broad ring segments correspond to the less andmore deformed parts of the sample. Similar characteristics ofDebye–Scherrer rings were observed for the Ti–W foam.Because the width of the sharp spots remained very narroweven after compression, these parts of the Debye–Scherrerrings were not evaluated. However, the continuous and broadsegments of the rings were cut to determine dislocationdensity and crystallite size. The individual peaks as a

Page 7: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

Fig. 6 – Debye–Scherrer diffraction rings of reflection 110 for (a) the as-processed state of the pure Ti foam and (b) aftercompression up to a plastic strain of �0.26. (c) Intensity distribution in direction x in figure (b) after subtracting thebackground. The red zone corresponds to the homogeneous and broad ring segments in (b) scattered from highly deformedparts of the sample.

Fig. 5 – Evolution of misorientation distribution during compression for (a) Ti and (b) Ti–W foams in an angle range of 21–61corresponding to approximately 95% of the misorientation distribution.

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6 413

function of the diffraction angle (2Θ) were obtained byintegrating the intensity along the rings (in direction x inFig. 6b), and these peaks were then used to construct the X-ray diffraction patterns for CMWP fitting. The average crystal-lite size values in the more deformed parts of the pure Ti andTi-W foams were 60 and 63 nm, respectively, as obtainedfrom CMWP fitting. It is noted that these crystallite sizevalues are much smaller than the grain sizes observed inthe EBSD images. This apparent discrepancy can be explainedby the breaking of X-ray coherency due to the small misor-ientations inside the grains (Gubicza, 2014). Therefore, XLPAmeasures the size of the subgrains or dislocation cells ratherthan the true grain size. The dislocation densities in the moredeformed parts of the Ti and Ti–W foams were (6.270.7)�1014 m�2 and (9.371.0)�1014 m�2, respectively. These valuescharacterize only the strongly deformed parts of the material.Therefore, the average dislocation density of the whole

material was calculated by considering the fraction of theless deformed parts in the foams. The latter quantity wasdetermined as the fraction of the intensities of the sharpspots in the entire Debye–Scherrer ring. The intensity wassummed in direction 2Θ (see Fig. 6b), and these integratedintensity values were plotted as a function of the coordinatex, as shown in Fig. 6c for reflection 110 of the pure Ti foamcompressed up to a plastic strain of �0.26. Thereafter, aspline was fitted to the parts of the intensity distribution thatare free of sharp intensity peaks (see Fig. 6c). The area underthis spline corresponds to the intensity scattered from themore deformed parts of the sample. The area under the sharppeaks but above the spline denotes the intensity scatteredfrom the less deformed volumes. The fractions of the moredeformed parts for the Ti and Ti–W foams were 55% and 63%,respectively. As the dislocation density is negligible in theless deformed volumes (it is lower than the detection limit,

Page 8: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6414

1013 m�2), the average dislocation density for the wholesample could be determined from the product of the volumefraction and the dislocation density of the more deformedparts. Therefore, the average dislocation density for theentire Ti and Ti–W foam materials (without pores) were(3.470.4)�1014 m�2 and (5.970.6)� 1014 m�2, respectively.The higher dislocation density in the Ti–W foam is attributedto the pinning effect of the solute W atoms on dislocationsformed inside the grains during compression. To the best ofour knowledge, this is the first report on the dislocationdensity in compressed Ti foams prepared by freeze castingfor biomedical applications.

Additionally, XLPA analysis was also employed for thedetermination of the fractions of oa4, oc4, and ocþa4

dislocations. The procedure was described in detail by Máthiset al. (2004). The fractions of oa4, oc4, and ocþa4

dislocations for both the Ti and Ti–W foams compressed upto a plastic strain of �0.26 were 6874%, 573%, and 2773%,respectively. Similar dislocation type fractions were obtainedin plastically deformed bulk Ti (Gubicza et al., 2008). Theabundance of oa4-type dislocations can be attributed totheir lowest energy due to the smallest Burgers vector.

4. Discussion

The measured elastic moduli of Ti and Ti–W foams were�23 GPa and �22 GPa, respectively, which are much lowerthan that of bulk Ti (approximately 114 GPa) owing to theuniformly distributed porosity. The relationship between theelastic moduli of the foam (E) and bulk Ti (Eb), and theporosity (P) can be expressed as follows (Luo and Stevens,1999):

E¼ Ebexpð�kPÞ; ð2Þ

where the value of parameter k depends on the material. Forthe foams used in this study, k¼0.05. In order to confirm thevalidity of Eq. (2) for freeze-cast Ti foams, an additional pureTi sample was synthesized with a porosity of 55%, asdetermined from the EBSD images. The higher porosity wasachieved by a lower sintering temperature of 1020 1C. For thissample, the elastic modulus and yield strength obtained bycompression were �7 GPa and �44 MPa, respectively. UsingEq. (2), the same value of k was obtained for this sample as forthe foams with 32%–33% porosity.

One of the most important mechanical parameters forsurgical implant materials is Young's modulus. In the case oftraditional implant materials (such as bulk stainless steel orTi), the elastic modulus of the implant (100–200 GPa) issignificantly higher than that of cortical bones [19–20 GPa(Rho et al., 1993)]. Therefore, the implant eventually bears themajority of the external load owing to its higher elasticmodulus. This stress-shielding effect typically leads to osteo-porosis of the bones. However, a porous structure in implantmaterials can reduce their elastic moduli to an appropriatevalue. The elastic moduli of the present Ti and Ti–W foams(�23 GPa) are similar to that of cortical bones, thus eliminat-ing stress-shielding effect. Therefore, these Ti foams arepromising candidates as bone implant materials.

Porosity also affects the plastic properties, i.e., yieldstrength and the strain hardening behavior of Ti foams.These parameters can be compared to the correspondingvalues for bulk Ti with Grade 2 purity with a similar grain sizeof �20 mm (Gubicza et al., 2008). Due to the porosity of pure Tifoam (33%), its yield strength (�196 MPa) is lower by a factorof 1.7 when compared with that obtained for bulk Ti (�330MPa). At the end of the compression test (at a plastic strain of�0.26), the flow stress of the Ti foam is primarily determinedby the dislocation density in the grain interiors and thevolume fraction of pores between the grains. The dislocationstrengthening can be taken into account using the followingTaylor equation:

sdisl ¼ αMTGbffiffiffiρ

p; ð3Þ

where α is a constant [�0.5 for Ti (Okazaki et al., 1977; Liu andMishnaevsky, 2014)], and MT is the Taylor factor [�3.05 fortexture-free Ti (Liu and Mishnaevsky, 2014; Panda et al.,2014)]. G is the shear modulus (�43 GPa for Ti), and b and ρ

are the Burgers vector of dislocations (�0.37 nm if thepopulation of the dislocation slip systems is taken intoaccount; see above) and dislocation density, respectively. Bysubstituting the dislocation density obtained via XLPA intoEq. (3), dislocation strengthening was calculated to be �455MPa. Because the porosity decreased minimally during thepresent compression test, its effect was taken into account bydividing the value obtained from Eq. (3) by a factor of 1.7,which was determined from the ratio of the yield strengthvalues of the bulk Ti and Ti foam. This stress value was thenadded to the yield strength, and a value of �464 MPa wasobtained, which is in good agreement with the experimentalflow stress value (�472 MPa). In the case of the Ti–W foam,dislocation strengthening was greater [�600 MPa as obtainedfrom Eq. (3)] at a similar plastic strain value (�0.26) owing tothe higher dislocation density. Taking the porosity effect intoaccount, the flow stress value of the Ti–W foam was 580 MPa,which is similar to the measured value (�525 MPa), consider-ing the experimental error. Therefore, we conclude that theaddition of W to the Ti foam increased the flow stress (i) byincreasing the threshold stress of plasticity as reflected by thehigher yield strength and (ii) by increasing the dislocationdensity by hindering the annihilation of dislocations. Thisstudy revealed that the addition of 5 wt% W to Ti foamincreased the compressive strength considerably while itsYoung's modulus remained similar to that of cortical bones,thereby improving the mechanical performance of the Tifoam for its potentially promising use in biomedicalapplications.

5. Conclusions

In this study, we analyzed the mechanical properties andmicrostructural evolution during compression of Ti and Ti–Wfoams with the porosity of 32–33% synthesized via freezecasting. The following conclusions were drawn:

1. The Young's moduli for both Ti and Ti–W foams were�23 GPa, which is in accordance with the predicted valuefrom the porosity and elastic modulus values of bulk Ti. On

Page 9: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6 415

the other hand, the yield strength was significantlyaffected by W content. The yield strength values for Tiand Ti–W foams were �196 MPa and �235 MPa, respec-tively. The higher yield strength for the Ti–W foam can beattributed to the solute hardening of tungsten. TheYoung's moduli of the foams are similar to the elasticmodulus of cortical bones, thereby eliminating theosteoporosis-causing stress-shielding effect and makingthe Ti foams potential candidates for use as surgicalimplant materials. It was also found that W additionincreased the compressive strength of Ti foam consider-ably without changing its beneficial low value of Young'smodulus for its potential use in biomedical applications.

2. During compression up to a plastic strain of �0.26, thedislocation densities in Ti and Ti–W foams increased to�3.4�1014 m�2 and �5.9�1014 m�2, respectively. Thehigher dislocation density in the Ti–W foam can beattributed to the pinning effect of the solute W atoms ondislocations in Ti grains. According to the best of ourknowledge, this is the first report on the dislocationdensity in compressed Ti foams prepared by freeze-casting for biomedical applications. It was revealed thatthe increase in flow stress during compression was causedby the increase of the dislocation density in the graininteriors.

3. The strain hardening behavior of the foams was investi-gated using the Kocks–Mecking plot. Strain hardeningdecreased gradually with increasing stress for both thefoams and the reference bulk Ti material. At each stressvalue, the strain hardening rate was significantly greaterfor the Ti–W foam than for the pure Ti foam, most probablydue to the higher rate of increase in dislocation densityduring compression, which was caused by the pinningeffect of the solute tungsten atoms on dislocations.

Acknowledgements

This work was supported by the Hungarian ScientificResearch Fund, OTKA, Grant No. K-109021. The authors aregrateful to Mr. Gábor Varga for the EBSD investigations andDr. György Krállics for providing the compressive stress-strain data for bulk Ti. Choi and Choe also would like toacknowledge supports from the National Research Founda-tion (NRF) of Korea (2015R1D1A1A01060773; 2009-0093814;2014R1A2A1A11052513; 2013K1A3A1A39074064).

r e f e r e n c e s

Choe, H., Abkowitz, S.M., Abkowitz, S., Dunand, D.C., 2005. Effectof tungsten additions on the mechanical properties of Ti-6Al-4V. Mater. Sci. Eng. A 396, 99–106, http://dx.doi.org/10.1016/j.msea.2005.01.051.

Choi, M., Hong, E., So, J., Song, S., Kim, B.-S., Yamamoto, A., Kim,Y.-S., Cho, J., Choe, H., 2014. Tribological properties of bio-compatible Ti–10W and Ti–7.5TiC–7.5W. J. Mech. Behav.Biomed. Mater. 30, 214–222, http://dx.doi.org/10.1016/j.jmbbm.2013.11.014.

D’Angeli, V., Belvedere, C., Ortolani, M., Giannini, S., Leardini, A.,2013. Load along the femur shaft during activities of daily

living. J. Biomech. 46, 2002–2010, http://dx.doi.org/10.1016/j.

jbiomech.2013.06.012.Dragomir, C.I., Ungar, T., 2002. Contrast factors of dislocations in

the hexagonal crystal system. J. Appl. Crystallogr. 35, 556–564,

http://dx.doi.org/10.1107/S0021889802009536.Erk, A.K., Dunand, D., Shull, K.R., 2008. Titanium with control-

lable pore fractions by thermoreversible gelcasting of TiH2.

Acta Mater. 56, 5147–5157, http://dx.doi.org/10.1016/j.

actamat.2008.06.035.Frary, M., Abkowitz, S., Abkowitz, S.M., Dunand, D.C., 2003.

Microstructure and mechanical properties of Ti/W and Ti-6Al-

4V/W composites fabricated by powder-metallurgy. Mater. Sci.

Eng. A 344, 103–112, http://dx.doi.org/10.1016/S0921-5093(02)

00426-4.Geetha, M., Singh, A.K., Asokamani, R., Gogia, A.K., 2009. Ti based

biomaterials, the ultimate choice for orthopedic implants – a

review. Prog. Mater. Sci. 54, 397–425, http://dx.doi.org/10.1016/

j.pmatsci.2008.06.004.Gubicza, J., 2014. X-ray Line Profile Analysis in Materials Science.

IGI-Global, Hershey, PA, USAhttp://dx.doi.org/10.4018/

978-1-4666-5852-3.Gubicza, J., Fogarassy, Zs, Krallics, Gy, Labar, J., Torkoly, T., 2008.

Microstructure and mechanical behavior of ultrafine-grained

titanium. Mater. Sci. Forum 589, 99–104, http://dx.doi.org/

10.4028/www.scientific.net/MSF.589.Hong, E., Ahn, B.Y., Shoji, D., Lewis, J.A., Dunand, D.C., 2011.

Microstructure and mechanical properties of reticulated tita-

nium scrolls. Adv. Eng. Mater. 13 (12), 1122–1127, http://dx.doi.

org/10.1002/adem.201100082.Ibrahim, A., Zhang, F., Otterstein, E., Burkel, E., 2011. Processing of

porous Ti and Ti5Mn foams by spark plasma sintering. Mater.

Des. 32 (1), 146–153, http://dx.doi.org/10.1016/j.

matdes.2010.06.019.Krishna, B.V., Bose, S., Bandyopadhyay, A., 2007. Low stiffness

porous Ti structures for load-bearing implants. Acta Biomater.

3, 997–1006, http://dx.doi.org/10.1016/j.actbio.2007.03.008.Kuzel, R., Klimanek, P., 1988. X-ray diffraction line broadening

due to dislocations in non-cubic materials. II. The case of

elastic anisotropy applied to hexagonal crystals. J. Appl.

Crystallogr. 21, 363–368, http://dx.doi.org/10.1107/

S002188988800336X.Lademo, O.-G., Hopperstad, O.S., Langseth, M., 1999. An evalua-

tion of yield criteria and flow rules for aluminium alloys. Int. J.

Plast. 15, 191–208, http://dx.doi.org/10.1016/S0749-6419(98)

00064-3.Li, H., Oppenheimer, S.M., Stupp, S.I., Dunand, D.C., Brinson, L.C.,

2004. Effect of pore morphology and bone ingrowth on

mechanical properties of microporous titanium as an ortho-

pedic implant material. Mater. Trans. 45 (4), 1124–1131, http:

//dx.doi.org/10.2320/matertrans.45.1124.Liu, H., Mishnaevsky Jr., L., 2014. Gradient ultrafine-grained

titanium: Computational study of mechanical and damage

behaviour. Acta Mater. 71, 220–233, http://dx.doi.org/10.1016/j.

actamat.2014.03.017.Long, M., Rack, H.J., 1998. Titanium alloys in total joint replace-

ment—a materials science perspective. Biomaterials 19,

1621–1639, http://dx.doi.org/10.1016/S0142-9612(97)00146-4.Luo, J., Stevens, R., 1999. Porosity-dependence of elastic moduli

and hardness of 3Y-TZP ceramics. Ceram. Int. 25, 281–286,

http://dx.doi.org/10.1016/S0272-8842(98)00037-6.Magee, D.J., 2008. In: Orthopedic Physical Assessment 6th ed.

Elsevier, Amsterdam.Mathis, K., Nyilas, K., Axt, A., Dragomir-Cernatescu, I., Ungar, T.,

Lukac, P., 2004. The evolution of non-basal dislocations as a

function of deformation temperature in pure magnesium

determined by X-ray diffraction. Acta Mater. 52, 2889–2894,

http://dx.doi.org/10.1016/j.actamat.2004.02.034.

Page 10: Mechanical behavior and microstructure of compressed Ti foams …heeman/paper/Mechanical... · 2019-09-22 · Mechanical behavior and microstructure of compressed Ti foams synthesized

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 a v i o r o f b i o m e d i c a l m a t e r i a l s 6 3 ( 2 0 1 6 ) 4 0 7 – 4 1 6416

Munoz, S., Pavon, J., Rodrıguez-Ortiz, J.A., Civantos, A., Allain, J.P.,

Torres, Y., 2015. On the influence of space holder in the

development of porous titanium implants: mechanical, com-

putational and biological evaluation. Mater. Charact. 108,

68–78, http://dx.doi.org/10.1016/j.matchar.2015.08.019.Niinomi, M., 2008. Mechanical biocompatibilities of titanium

alloy for biomedical applications. J. Mech. Behav. Biomed.

Mater. 1, 30–42, http://dx.doi.org/10.1016/j.jmbbm.2007.07.001.Niu, W., Bai, C., Qiu, G., Wang, Q., 2009. Processing and properties

of porous titanium using space holder technique. Mater. Sci.

Eng. A 506 (1–2), 148–151, http://dx.doi.org/10.1016/j.

msea.2008.11.022.Okazaki, K., Odawara, T., Conrad, H., 1977. Deformation

mechanism map for titanium. Scr. Metall. 11, 437–440, http:

//dx.doi.org/10.1016/0036-9748(77)90151-X.Panda, S., Sahoo, S.K., Dash, A., Bagwan, M., Kumar, G., Mishra, S.

C., Suwas, S., 2014. Orientation dependent mechanical prop-

erties of commercially pure (cp) titanium. Mater. Charact. 98,

93–101, http://dx.doi.org/10.1016/j.matchar.2014.10.011.Rho, J.Y., Ashman, R.B., Turner, C.H., 1993. Young’s modulus of

trabecular and cortical bone material: ultrasonic and micro-

tensile measurements. J. Biomech. 26, 111–119, http://dx.doi.

org/10.1016/0021-9290(93)90042-D.Ribarik, G., Gubicza, J., Ungar, T., 2004. Correlation between

strength and microstructure of ball-milled Al–Mg alloys

determined by X-ray diffraction. Mater. Sci. Eng. A 387–389,

343–347, http://dx.doi.org/10.1016/j.msea.2004.01.089.

Ryan, G., Pandit, A., Apatsidis, D.P., 2006. Fabrication methods ofporousmetals for use in orthopedic applications. Biomaterials 27,2651–2670, http://dx.doi.org/10.1016/j.biomaterials.2005.12.002.

Sauer, B.W., Weinstein, A.M., Klawitter, J.J., Hulbert, S.F., Leonard,R.B., Bagwell, J.G., 1974. The role of porous polymeric materialsin prosthesis attachment. J. Biomed. Mater. Res. 8 (3), 145–153,http://dx.doi.org/10.1002/jbm.820080315.

Siegkas, P., Petrinic, N., Tagarielli, V.L., 2016. Measurements andmicro-mechanical modelling of the response of sinteredtitanium foams. J. Mech. Behav. Biomed. Mater. 57, 365–375,http://dx.doi.org/10.1016/j.jmbbm.2016.02.024.

Taniguchi, N., Fujibayashi, S., Takemoto, M., Sasaki, K., Otsuki, B.,Nakamura, T., Matsushita, T., Kokubo, T., Matsuda, S., 2016.Effect of pore size on bone ingrowth into porous titaniumimplants fabricated by additive manufacturing: an in vivoexperiment. Mater. Sci. Eng. C 59, 690–701, http://dx.doi.org/10.1016/j.msec.2015.10.069.

Yamamoto, D., Kawai, I., Kuroda, K., Ichino, R., Okido, M., Seki, A.,2012. Osteocondustivity and hydrophilicity of TiO2 coatings onTi substrates prepared by different oxidizing process. Bioinorg.Chem. Appl. 495218, 1–7, http://dx.doi.org/10.1155/2012/495218.

Zhao, S., Li, S.J., Hou, W.T., Hao, Y.L., Yang, R., Misra, R.D.K., 2016.The influence of cell morphology on the compressive fatiguebehavior of Ti-6Al-4V meshes fabricated by electron beammelting. J. Mech. Behav. Biomed. Mater. 59, 251–264, http://dx.doi.org/10.1016/j.jmbbm.2016.01.034.