Sie sind auf Seite 1von 9

Acta Biomaterialia 3 (2007) 359367 www.elsevier.

com/locate/actabiomat

Microstructure and deformation behavior of biocompatible TiO2 nanotubes on titanium substrate q


G.A. Crawford a, N. Chawla
b

a,*

, K. Das b, S. Bose b, A. Bandyopadhyay

a School of Materials, Arizona State University, Tempe, AZ 85287-8706, USA School of Mechanical and Materials Engineering, Washington State University, Pullman, WA 99164, USA

Received 20 April 2006; received in revised form 7 August 2006; accepted 16 August 2006

Abstract Titanium oxide coatings have been shown to exhibit desirable properties as biocompatible coatings. We report on the quantitative microstructure characterization and deformation behavior of TiO2 nanotubes on Ti substrate. Nanotubes were processed using anodic oxidation of Ti in a NaF electrolyte solution. Characterization of the as-processed coatings was conducted using scanning electron microscopy and focused ion beam milling. Increases in anodization time had no signicant eect on tube diameter or tube wall thickness. Coating thickness, however, increased with time up to 2 h of anodization, at which point an equilibrium thickness was established. Nanoindentation was used to probe the mechanical response in terms of Youngs modulus and hardness. Progressively higher values of elastic modulus were obtained for thinner lms consistent with increasing eects of the Ti substrate. A possible deformation mechanism of densication of the porous oxide and wear of the dense surface is suggested and discussed. 2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Nanoindentation; Biocompatible; TiO2; Porosity; Coating

1. Introduction Titanium (Ti) and Ti alloys have been used extensively as bone-implant materials due to their high strength-toweight ratio, good biocompatibility and excellent corrosion resistance [1,2]. Titanium, however, being relatively inert, cannot directly bond to bone, and osseointegration via the natural oxide (TiO2) is a long process [3,4]. Hence, there is an increased interest in reducing the time needed for osseointegration. In an eort to enhance the cellimplant material interaction and increase lifetime, bioactive ceramic based coatings have been applied to Ti implants, most notably hydroxyapatite (HA) [58]. Recently, TiO2 has been suggested as a potential alternative to HA coatings. The advantage of using TiO2 is that
q Research presented at the TMS 2006 Biological Materials Science Symposium. * Corresponding author. E-mail address: nchawla@asu.edu (N. Chawla).

it can be grown directly on the Ti surface, by cost-eective techniques such as anodic oxidation [911]. Also, it is well known that one problem with HA coatings is poor adhesion strength at the HA/Ti interface [58]. Using anodic oxidation, TiO2 is formed with a chemical bond between the oxide and Ti substrate that likely results in enhanced adhesion strength. Indeed, a bone-like apatite layer is formed on TiO2 in simulated body uid (SBF) [9,10]. Furthermore, researchers have suggested that TiO2 with a 3-D micro/nanoporous structure may enhance apatite formability when compared to dense TiO2 [11]. Recently, Gong et al. [12] have shown that TiO2 coatings, consisting of regular self-assembled nanotube arrays, may be formed on the surface of Ti by anodic oxidation. Treating the TiO2 nanotubes with NaOH solution induces the growth of nanosized HA when subjected to SBF [13]. Furthermore, nanophase ceramics such as HA, Al2O3 and TiO2 enhance long-term osteoblast functions relative to typical microstructure forms [14]. Thus, the apatite-forming and bone-bonding ability of titanium bone-implant materials may be

1742-7061/$ - see front matter 2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actbio.2006.08.004

360

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

signicantly enhanced by anodically oxidized nanotubes on the surface of Ti. One of the most important attributes of materials in medical and dental applications is its mechanical properties [1,2]. This is particularly important in bone-implant materials, where long-term in vivo structural stability is crucial. Here, bone resorption occurs due to the mismatch in mechanical properties (i.e., elastic modulus) between bone and the implant material, causing implant loosening and eventual failure [15]. Thus, it is important to characterize the mechanical behavior of these implant materials. Furthermore, it is important to establish the relationship between microstructure and mechanical behavior, in an eort to elucidate deformation mechanisms. Mechanical characterization of thin coatings is challenging because of the small length scales involved. In this regard, nanoindentation is a very appropriate technique for these materials because of its low load (1 lN) and small displacement (1 nm) resolution [16]. Nanoindentation is an eective technique for probing mechanical properties of thin (<10 lm) coated systems including soft coatings on hard substrates [17,18], hard coatings on soft substrates [1921] and multilayered systems [19,22,23]. This technique has also been used to investigate the mechanical response of TiO2 with nanophase microstructure [24,25] and of thin TiO2 coatings on Ti substrates [20,21]. To our knowledge, a systematic investigation of the microstructure and mechanical behavior of self-assembled, vertically aligned nanotubes has not yet been reported. Thus, in this study we focus on the relationship between microstructure characterization and deformation behavior, by nanoindentation, of TiO2 nanotubes on Ti substrate. We have studied the eect of changes in anodic oxidation time on tube diameter, tube wall thickness and coating thickness. Nanoindentation was conducted to probe the mechanical properties, such as Youngs modulus and hardness. Post-indentation characterization was employed to provide insight into deformation mechanisms in this system. 2. Materials and experimental procedure Thin titanium sheets (99.8% purity, Supra Alloys, CA), 0.5 mm in thickness, were used for fabrication of TiO2 nanotubes. Disks (12 mm diameter) were ground with SiC paper and ultrasonically cleaned in distilled water. Final polishing of the sample was performed with 1 lm alumina suspension. A two-electrode electrochemical anodization cell, with a platinum cathode and Ti anode, was used to fabricate the nanotube arrays. Anodization voltage was kept constant at 20 V with a DC power supply (HewlettPackard 060 V/050 A, 1000 W). The electrolyte was made by dissolving sodium uoride (NaF, SigmaAldrich), citric acid (J.T. Baker), 1 M sulfuric acid (Fischer Scientic) in a ratio so that the nal electrolyte component concentrations were F 0.1 mol/l, SO2 4 1.0 mol/l, citric acid 0.2 mol/l. The electrolyte pH was

adjusted to 4.5 using NaOH solution. All experiments were performed at room temperature. Characterization of coating surfaces and cross sections was carried out using a eld emission scanning electron microscope (FESEM, Hitachi S4700). The inner diameter and wall thickness of the tubes were measured by segmenting SEM micrographs and using image analysis software (ImageJ, Gaithersburg, MD). Micrographs for quantitative analysis were taken from at least two dierent locations on each sample. This was done to ensure that the measurements were representative of the overall microstructure and to gauge the degree of sample variability. About 500 measurements of tube diameter were taken per sample location. Tube wall thicknesses were measured at random from an average of 50 tubes per sample location. Characterization of coating/substrate cross sections was carried out by fracturing the coating by bending, and examining the detached TiO2 coating. SEM was used to examine both the bottom of the coating as well as the mating Ti substrate. Coating thickness was measured directly from SEM micrographs. At least six measurements were taken from dierent areas of the cross section to ensure that the measurements reected the actual coating thickness. The coating cross section was also studied by a relatively nondestructive means. A focused ion beam (FIB) with high-resolution SEM (FEI Nova 200 SEM/FIB, Hillsboro, OR) was used for ion milling and characterization of the TiO2/Ti cross section. Ion milling was conducted at 30 keV, with a milling and cleaning cross section current of 0.4 nA and 10 pA, respectively. In order to minimize ion beam damage, a protective platinum layer ($1 lm thick) was deposited in situ using the electron source, prior to ion milling. The mechanical properties of the coatings were probed by nanoindentation. A commercial nanoindenter (MTSNano Instruments XP-II, Minneapolis, MN) was used with a continuous stiness measurement (CSM) capability [26]. Here, a small-harmonic high-frequency amplitude is superimposed over indentation loading and the contact stiness of the sample is measured from the displacement response at the excitation frequency. From this contact stiness, the Youngs modulus of the material can be derived. Thus, the modulus or hardness can be determined instantaneously as a function of depth. For all experiments, indentation was carried out using a Berkovich (three-sided pyramid) indenter. Calibration for load and hardness was performed on a reference sample of fused silica. Indentation experiments were conducted in displacement control to depths of 2000 nm for each sample. At least ve indentations were obtained for each condition. 3. Results and discussion 3.1. Microstructure characterization We begin with the characterization of the surface and cross section of the TiO2 coatings. Fig. 1 shows a SEM

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

361

Fig. 1. FESEM micrograph of TiO2 nanotubes formed by anodic oxidation of Ti in NaF-containing electrolyte (0.1 mol/l F), at pH 4.5, and constant potential of 20 V for 2 h. Table 1 Dimensions of TiO2 nanotubes fabricated through anodic oxidation of Ti in NaF-containing electrolytes (0.1 mol/l F), at pH 4.5, and constant potential of 20 V Anodization time (h) Tube dimensions (nm) Inner diameter 0.25 0.5 2 4 45 10 43 8 58 12 50 9 Wall thickness 12 2 12 2 16 2 15 2 Tube length 234 6 245 14 650 20 625 23

micrograph of Ti anodized for 2 h in NaF-containing electrolyte (0.1 mol/l F), at a pH of 4.5, and constant potential of 20 V. A uniform structure of thin-walled TiO2

nanotubes was obtained. Table 1 summarizes the quantitative measurements of the important dimensions of the tubes. Tube diameter, tube wall thickness and tube length were 3570, 1018 and 230670 nm, respectively. More importantly, tube diameter and wall thickness did not appear to vary with anodization time, for the range of times examined here. Our results are consistent with those found by other researchers under similar processing conditions [29]. A small increase in tube diameter and wall thickness with increasing anodization time, from 30 min to 2 h, is observed. Given the standard deviation of the reported data, however, this dierence is not signicant. Fig. 2 shows a cross section of the fractured TiO2 coating. The mating surfaces of the bottom of the coating and the top of the substrate are also shown. Slight variations in the tube wall thickness, or ridges, at regular spaced intervals are observed. The possible origin of these ridges is discussed below. The bottom surface of the TiO2 nanotube coating is characterized by a series of regularly spaced bumps (Fig. 2b). These bumps represent the pore tip of each individual nanotube. The bottom surface clearly shows the bottom of the barrier layer, which acts as a thin barrier between the Ti substrate and the tubular structure. The formation of this and other structures are discussed below. The mating Ti surface (Fig. 2c) shows corresponding concave dimples that match that of the TiO2 coating. This cup and cone type structure at the interface may encourage mechanical interlocking, and may increase the adhesion strength at the metal/ceramic interface. The cross section of the interface was also observed by micromilling using a focused ion beam (FIB). The FIB allows one to obtain a smooth surface at the coating/

Fig. 2. FESEM micrograph showing (a) fractured cross section, (b) bottom surface of TiO2 nanotubes, (c) surface of Ti substrate, following anodic oxidation of Ti in NaF-containing electrolytes (0.1 mol/l F), at pH 4.5, and constant potential of 20 V for 4 h.

362

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

substrate interface. This is particularly important for the materials studied here, because the substrate and coating have large dierences in stiness and hardness. As a result, conventional mechanical polishing would have resulted in preferential removal of the softer phase. Fig. 3 shows a SEM micrograph of the cross section of TiO2/Ti sample produced by the FIB milling technique. The platinum layer can be observed at the top of the micrograph. The interface between TiO2 and Ti is quite uniform. Once again, ridge markings can be seen on the walls of the TiO2 nanotubes. Using the SEM micrographs of the cross sections, quantitative measurements of the coating thickness were obtained. Fig. 4 shows a plot of coating thickness (tube length) vs. anodization time. It appears that there is a signicant increase in coating thickness with time between 30 min and 2 h. Between 2 and 4 h, however, the thickness is relatively unchanged. The origin of this behavior, as well as other microstructural observations, are now discussed. TiO2 nanotube formation in F-containing electrolytes is the result of two competing, electric eld-assisted processes. Hydrolysis of Ti metal to form TiO2 occurs by the following reaction [2729]: Ti4 + 2H2 O ! TiO2 + 4H 1

Fig. 4. Average coating thickness vs. anodization time for Ti samples anodized in NaF-containing electrolyte (0.1 mol/l F), at pH 4.5, and constant potential of 20 V.

In addition, chemical dissolution of TiO2 at the oxide/ electrolyte interface occurs, resulting in the formation of [TiF6]2 [2729]: TiO2 + 6HF ! [TiF6 ]2 + 2H2 O + 2H 2

The growth process of the tubes may be divided into three main processes: (i) initial barrier layer formation, (ii) formation of uniformly distributed pores and (iii) separation of interconnected pores into nanotubes. A detailed understanding of the exact mechanisms for ordered TiO2 nanotube formation is still not clear. Several possible

Fig. 3. FESEM micrograph of FIB milled TiO2/Ti cross section prepared by anodization of Ti in NaF-containing electrolytes (0.1 mol/l F), at pH 4.5, and constant potential of 20 V for 4 h. Insert shows surface morphology of sample.

mechanisms for formation of these tubes, however, have been proposed. During the onset of anodization, a barrier oxide layer begins to form, causing an exponential decrease in the anodic current density. In uorine-free solutions, both the thickness of the barrier layer and the measured current density reach steady state [27,2931]. In uorinecontaining solutions, however, a nanoporous structure begins to form due to the chemical dissolution of the barrier oxide layer (Eq. (2)). An increase in current density takes place during this process. Over time pores become assembled in an ordered fashion and become unconnected. Once an ordered nanotube structure is established, the current density again stabilizes, at a value greater than that in a uorine-free solution [27,30,31]. At this point, further anodic oxidation simply increases the length of the tubes but does not signicantly aect the tube structure or the anodic current density. The exact mechanism by which nanopores form and are separated into nanotubes is still the subject of some debate. We address the case of ordered nanopore formation rst. Several researchers have suggested that nanopore formation occurs by random local dissolution of the TiO2 surface which drives the formation of pores [3032]. Here pore formation is driven by localized dissolution of TiO2 which reduces the lm thickness locally, increasing the electric eld intensity at the bottom of the pore and inducing the formation of new oxide. Thus, the reactions given above (Eqs. (1) and (2)) take place simultaneously at the bottom of the pore and cause the pore to grow further into the Ti substrate [29,31]. Beranek et al. [30] suggest that pore formation takes place initially at random locations, and selfordering is merely a product of the competition between growing pores. Raja et al. [27], on the other hand, suggest that the ordering of pores may be a result of local surface perturbations. Here increases in strain energy of the oxide lm may be caused by variations in dielectric constant and electrostriction coecient throughout the lm. In this case,

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

363

uoride ions could migrate to regions of high strain energy density, causing the migration of hydrogen ions to maintain electrical neutrality, and leading to the dissolution of Ti4+ ions. As a result, mass ow may occur, such that low strain energy locations grow at the expense of the dissolving high strain energy areas (i.e., pores). Following pore formation, neighboring nanopores must be separated to form individual nanotubes such as those seen in Fig. 1. Mor et al. [31] suggested that the unanodized regions between pores are susceptible to eld-assisted oxidation/dissolution causing the formation of voids between nanopore regions. Thus, nanotubes are the result of simultaneous growth of voids and pores. However, Raja et al. [27] have proposed that void growth in equilibrium with pore growth is an unlikely mechanism for the formation of tubes with well dened circular cross sections. In contrast, they propose that the separation of pores into individual nanotubes may be a result of repulsive forces between cation vacancies. Negative cation vacancies are formed when Ti4+ ions dissolve into solution. It is postulated that these negatively charged vacancies migrate radially to the center of the interpore region, following the direction of the electric eld. Vacancies arriving at the interpore region, originating from neighboring pores, will have the same sign and will repel each other, keeping an equilibrium distance. If the dissolution rate at the Ti/ TiO2 interface is higher than the generation of oxygen vacancies (produced to maintain electrical neutrality), the cation vacancies may condense to form voids between the neighboring pores [27]. Let us now address the issue of ridge formation in the tubes (Figs. 1 and 3). Macak et al. [32] have suggested that these ridges or inconsistencies in wall thickness are a result of current oscillations during the anodization process. Each current transient is accompanied by a change in pH at the pore tip and, thus, a temporary increase in the dissolution rate. In fact, by decreasing the diusion constant of the electrolyte, they showed that changes in pH may be damped and variations in dissolution rate minimized, resulting in nanotubes with smooth tube walls [32]. Finally, the evolution of coating thickness can be explained as follows. An equilibrium thickness results when there is a balance between the rate of porous oxide formation at the pore tip and the chemical dissolution rate of the TiO2 at the pore mouth [27,2931]. Thus, with anodization times longer than 2 h, new oxide is being formed at the metal/ceramic interface. The thickness of the coating, however, is maintained by chemical dissolution at the ceramic/ electrolyte interface. In fact, it has been shown that by adjusting the electrolyte solution and using a voltage sweep technique, a pH gradient may be introduced, such that there is localized acidication at the pore tip but a basic solution at the pore walls and mouth [28,32]. The chemical dissolution rate at the top of the tube is impeded, while oxide formation at the pore tip is increased. Using this logic, nanotubes with lengths much longer than those reported here have been fabricated (i.e. 2.5 lm [28],

7.7 lm [32]). Long (i.e. 4.4 lm [29]) TiO2 nanotubes have also been formed using constant potential (rather than voltage sweeping) anodization. 3.2. Nanoindentation In this section we describe nanomechanical characterization of the TiO2 nanotube structures on Ti. Fig. 5a shows representative load vs. displacement curves for the samples with varying anodization time. Fig. 5b and c shows the extracted apparent Youngs modulus and hardness, respectively, obtained from the continuous stiness measurement (CSM) system. The term apparent is used because the measured modulus and hardness are not only a function of the coating, but are also inuenced by the substrate. This is immediately apparent from the modulus vs. displacement curves for the thinner coatings (15 and 30 min). Here, with increasing depth, the measured modulus approached that of the Ti substrate. Average coating thickness for the samples are designated by a dotted vertical line in the plots of apparent modulus vs. displacement (Fig. 5b). Note, because coating thicknesses are similar for the 2 and 4 h and also the 15 and 30 min, a single line is displayed for each set. Thus, from inspection of Fig. 5b, total nanoindentation depth is much larger than the thickness of the coating. In order to gain insight into the underlying deformation mechanisms, SEM of the indentations was conducted. Fig. 6a and b shows a SEM micrograph of typical nanoindentations for the 15 min and 2 h samples respectively. Note that the TiO2 coating anodized for 15 min shows evidence of delamination. Indeed, delamination was observed in almost all indentations in the 15 and 30 min samples. We believe that coating delamination occurred on unloading, due to residual stresses induced from the loading cycle. This conclusion can be made because there were no consistent uctuations in the experimental data that suggest evidence of delamination during loading. None of the indentations for 2 or 4 h, however, showed signs of delamination. Thus, it is important to note that coating delamination occurred only in thin coatings (230250 nm) and did not occur in thicker coatings (600650 nm). This suggests that the adhesion strength may be enhanced with increases in coating thickness. The exact mechanism for this phenomenon is not well understood. However, increased thickness of the porous layer may retard the build-up of stress needed to cause delamination during the loading cycle. An additional characteristic of indentations in thicker coatings were wear markings on the indentation surface (Fig. 6b). These markings suggest some degree of frictional sliding between the indenter and the TiO2 surface. We also observe an increase in density of the coating, which decreased radially from tip of the indenter. This suggests the densication of nanotubes under nanoindentation. The exact mechanism of densication is not clear. It is likely, however, that fracture of the nanotubes accompanies densication.

364

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

Fig. 5. Characteristic (a) load, (b) apparent Youngs modulus and (c) apparent hardness vs. nanoindentation displacement (depth) for Ti samples anodized in NaF-containing electrolytes (0.1 mol/l F), at pH 4.5, and constant potential of 20 V for various times.

Fig. 6. FESEM micrographs of typical nanoindentations on the surface of Ti anodized for (a) 15 min and (b) 2 h in NaF-containing electrolytes (0.1 mol/l F), at pH 4.5, and constant potential of 20 V.

From the SEM micrographs and the indentation data, we can construct a proposed model for deformation in these systems. There appear to be two main deformation processes: (i) densication of the porous nanotube structure and (ii) wear of the dense surface. Fig. 7 shows a schematic of the proposed deformation process. Regions immediately under the indenter tip are subject to densication, while those at the sides of the indenter are under shear

stresses that induce both densication and wear of the TiO2 nanotubes. The deformation mechanisms can also be correlated to the measured modulus vs. depth curves. Three distinct regions may be dened and are shown in Fig. 8. Region I is characterized by a linear increase in modulus with increasing time. In this region, the increasing modulus is primarily due to increased densication of the TiO2

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

365

Fig. 7. Schematic illustration of the proposed deformation process during nanoindentation of TiO2 nanotubes.

Fig. 8. Typical apparent Youngs modulus vs. displacement plot indicating three characteristic regions of the deformation process. Dotted vertical line represents the average coating thickness.

Since densication plays a major role in the deformation process of TiO2 nanotubes, it would be logical for there to be a strong relationship between coating thickness and mechanical response. Fig. 9 shows a plot of composite modulus vs. coating thickness for all samples. Here composite modulus is the modulus taken from the plateau in Region III (Fig. 8) of the nanoindentation data. First, it is important to point out that for all cases the composite modulus is less than the modulus of the underlying substrate itself. Thus, the modulus of the coating is most likely lower that that of the Ti substrate. Also, in this plot we can clearly see a decrease in composite modulus with increasing thickness. This can be attributed to a larger contribution of the more compliant coating, relative to the Ti substrate. Thus, progressively larger values of composite modulus for decreases in coating thickness suggest an increasing substrate eect in thinner coatings.

nanotubes. Region II is characterized by a parabolic increase in the indentation modulus. In this region the indentation depth is comparable to the thickness of the TiO2 coating, i.e. the indenter is approaching the substrate surface. Thus, increases in modulus are the result of an increasing contribution from the substrate. In addition, there are minor contributions of densication of the outer regions of the indentation and wear between the indenter surface and the dense TiO2 surface. Note, as the TiO2 coating under the indenter becomes increasingly dense, the modulus increases. Finally, Region III describes the region where the coating has become nearly fully dense. This region is characterized by a plateau in indentation modulus, which we now term the composite modulus, as it is a combination of moduli of the dense coating and the Ti substrate. In this region, minor contributions of densication of the outer indentation regions and wear of the dense TiO2 surface take place.

Fig. 9. Plot of composite modulus vs. coating thickness, from Region III (Fig. 8) of the experimental data.

366

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367

From the above analysis, it is dicult to extract a precise value of the Youngs modulus of the porous coating itself. In nanoindentation of thin lms, it is common to determine the properties of the lm at indentation depths less than 10% of the lm thickness [33]. This is to minimize contributions from the substrate. In our case, 10% of the lm thickness is of the order of 60 nm or less. The roughness of the coating is higher than this indentation depth. Therefore, it is dicult to determine the Youngs modulus of the porous coating directly from our indentation data. It is well known, however, that the modulus of ceramic materials decreases with increases in porosity. Several researchers have proposed theoretical models to predict the eect of porosity on the elastic properties of ceramic materials [3436]. One of the simplest models is the rule of mixtures, commonly used for continuous ber reinforced composites. It assumes an isostrain condition (longitudinal loading, parallel to the bers) or isostress condition (transverse loading, perpendicular to the bers). In the longitudinal orientation, the Youngs modulus of the composite, Ec, is given by: E c Ef V f E m V m ; where E and V correspond to Youngs modulus and volume fraction, respectively. The subscripts f and m denote ber and matrix. Hashin and Rosen [37] derived a relationship for a transversely isotropic composite consisting of cylindrical bers in various aligned packing arrangements. By assuming the bers to be cylindrical pores (i.e., Youngs modulus of zero) their equations also reduced to a rule of mixtures type of relationship: Ec Em 1 P ; where Em is the matrix material and P is the fraction of porosity. In our system, the nanotubes are aligned, as is the porosity between the tubes. Thus, we can use the rule of mixtures to determine the Youngs modulus of this porous system. For our analysis we assume the volume fraction of pores in the coating is approximately 72% (from image analysis) and a modulus for dense amorphous TiO2 of 130150 GPa (for the tubes themselves) [33]. This yields a Youngs modulus of TiO2 nanotubes of approximately 3643 GPa. These results seem reasonable, since our nanoindentation results indicate that the modulus of the coating is signicantly lower than that of the Ti substrate. 4. Conclusions We have reported on the quantitative microstructure characterization and deformation behavior of TiO2 nanotubes on Ti substrate. The following conclusions can be made: 1. Nanotubes were processed using anodic oxidation, with NaF electrolyte. 2. Increases in anodization time showed no signicant eect on tube diameter or tube wall thickness.

3. Coating thickness increased up to 2 h of anodization, at which point an equilibrium thickness was established. Equilibrium thickness is achieved when the rate of chemical dissolution of TiO2 at the pore mouth matches that of new oxide formation at the pore tip. 4. Nanoindentation results yielded progressively higher values of elastic modulus for thinner lms, consistent with increasing eects of the Ti substrate. 5. A possible deformation mechanism of densication of the porous oxide and wear of the dense surface appears to be taking place. Acknowledgements The authors are grateful for nancial support from the Oce of Naval Research (Dr. A.K. Vasudevan, program manager) through a subcontract from Washington State University. The authors also acknowledge M. Koopman, K. Carlisle and Professor K.K. Chawla, at the University of Alabama at Birmingham, for their assistance with nanoindentation. References
[1] Katz JL. Application of materials in medicine and dentistry: orthopedic applications. In: Ratner BD, Homan AS, Schoen FJ, Lemons JE, editors. Biomaterials science: an introduction to materials in medicine. San Diego (CA): Academic Press; 1996. p. 33546. [2] Park JB, Lakes RS. Biomaterials: an introduction. New York: Plenum Press; 1992, p. 79115. [3] Branemark PI. Osseointegration and its experimental background. J Prosthet Dent 1983;59:399410. [4] Li P, Ducheyne P. Quasi-biological apatite lm induced by titanium in simulated body uid. J Biomed Mater Res 1998;41:3418. [5] Degroot K, Geesink R, Klein CPAT, Serekian P. Plasma sprayed coatings of hydroxyapatite. J Biomed Mater Res 1987;21:137581. [6] Mcpherson R, Gane N, Bastow TJ. Structural characterization of plasma-sprayed hydroxylapatite coatings. J Mater SciMater M 1995;6:32734. [7] Kurzweg H, Heimann RB, Troczynski T. Adhesion of thermally sprayed hydroxyapatite-bond-coat systems measured by a novel peel test. J Mater SciMater M 1998;9:916. [8] Lin CM, Yen SK. Characterization and bond strength of electrolytic HA/TiO2 double layers for orthopedic applications. J Mater Sci Mater M 2004;15:123746. [9] Kim HM, Kaneko H, Kawashita M, Kokubo T, Nakamura T. Mechanism of apatite formation on anodically oxidized titanium metal in simulated body uid. Key Eng Mater 2004;254 256:7414. [10] Kim HM, Himeno T, Kawashita M, Lee JH, Kokubo T, Nakamura T. Surface potential change in bioactive titanium metal during the process of apatite formation in simulated body uid. J Biomed Mater Res 2003;67A:13059. [11] Yang B, Uchida M, Kim HM, Zhang X, Kokubo T. Preparation of bioactive titanium metal via anodic oxidation treatment. Biomaterials 2004;25:100310. [12] Gong D, Grimes CA, Varghese OK, Hu W, Singh RS, Chen Z, et al. Titanium oxide nanotubes arrays prepared by anodic oxidation. J Mater Res 2001;16:33314. [13] Oh SH, Finones RR, Daraio C, Chen LH, Jin S. Growth of nanoscale hydroxyapatite using chemically treated titanium oxide nanotubes. Biomaterials 2005;4938:4943.

G.A. Crawford et al. / Acta Biomaterialia 3 (2007) 359367 [14] Webster TJ, Celaletdin E, Doremus RH, Siegel RW, Bizios R. Enhanced functions of osteoblasts on nanophase ceramics. Biomaterials 2000;21:180310. [15] Huiskes R, Weinans H, Vanrietbergen B. The relationship between stress shielding and bone-resorption around total hip stems and the eects of exible materials. Clin Orthop Relat Res 1992;274:12434. [16] Oliver WC, Pharr GM. An improved technique for determining hardness and elastic-modulus using load and displacement sensing indentation experiments. J Mater Res 1992;7:1564. [17] Tsui TY, Pharr GM. Substrate eects on nanoindentation mechanical property measurement of soft lms on hard substrates. J Mater Res 1999;14:292301. [18] Tsui TY, Vlassak J, Nix WD. Indentation plastic displacement eld: Part I. The case of soft lms on hard substrates. J Mater Res 1999;14:2196203. [19] Tsui TY, Vlassak J, Nix WD. Indentation plastic displacement eld: Part II. The case of hard lms on soft substrates. J Mater Res 1999;14:22049. [20] Bahr DF, Woodcock CL, Pang M, Weaver KD, Moody NR. Indentation induced lm fracture in hard lmsoft substrate systems. Int J Fract 2003;119:33949. [21] Pang M, Bahr DF. Thin-lm fracture during nanoindentation of a titanium oxide lmtitanium system. J Mater Res 2001;16:263443. [22] Deng X, Cleveland C, Karcher T, Koopman M, Chawla N, Chawla KK. Nanoindentation behavior of nanolayered metalceramic composites. J Mater Eng Perform 2005;14:41723. [23] Deng X, Chawla N, Chawla KK, Koopman M, Chu JP. Mechanical behavior of multilayered nanoscale metalceramic composites. Adv Eng Mater 2005;7:1099108. [24] Mayo MJ, Siegel RW, Narayanasamy A, Nix WD. Mechanical properties of nanophase TiO2 as determined by nanoindentation. J Mater Res 1990;5:107381. [25] Han Y, Hong SH, Xu KW. Porous nanocrystalline titania lms by plasma electrolytic oxidation. Surf Coat Technol 2002;154:3148.

367

[26] Oliver WC, Pharr GM. Measurement of hardness and elastic modulus by instrumented indentation: advances in understanding and renements in methodology. J Mater Res 2004;19:320. [27] Raja KS, Misra M, Paramguru K. Formation of self-ordered nanotubular structure of anodic oxide layer on titanium. Electrochim Acta 2005;51:15465. [28] Macak JM, Tsuchiya H, Schmuki P. High-aspect-ratio TiO2 nanotubes by anodization of titanium. Angew Chem Int Ed 2005;44:21002. [29] Cai Q, Paulose M, Varghese OK, Grimes CA. The eect of electrolyte composition on the fabrication of self-organized titanium oxide nanotube arrays by anodic oxidation. J Mater Res 2005;20:2306. [30] Beranek R, Hildebrand H, Schmuki P. Self-organized porous titanium oxide prepared in H2SO4/HF electrolytes. Electrochem Solid State Lett 2003;6:B124. [31] Mor GK, Varghese OK, Paulose M, Mukherjee N, Grimes CA. Fabrication of tapered, conical-shaped titania nanotubes. J Mater Res 2003;18:258893. [32] Macak JM, Tsuchiya H, Taveira L, Aldabergerova S, Schmuki P. Smooth anodic TiO2 nanotubes. Angew Chem Int Ed 2005;44:74635. [33] Fischer-Cripps AC. Nanoindentation. 2nd ed. New York: Springer; 2004. [34] Rice RW. Microstructure dependence of mechanical behavior of ceramics. In: MacCrone RK, editor. Treatise on materials science and technology, vol. 11. New York: Academic Press; 1977. p. 20331. [35] Chawla KK. Composite materials science and engineering. 2nd ed. New York: Springer; 1998, p. 30346. [36] Rice RW. Mechanical properties of ceramics and composites. New York: Marcel Dekker; 2000, p. 45861. [37] Hashin Z, Rosen BW. The elastic moduli of ber-reinforced materials. J Appl Mech 1964;31:22332.

Das könnte Ihnen auch gefallen