In situ imaging


Biomaterials
Volume 24, Issue 11 , May 2003, Pages 1837-1852

In situ imaging and impedance measurements of titanium surfaces using AFM and SPIS

Jane P. Bearingera, b, 1, Christine A. Ormec and Jeremy L. Gilbert, , d

a Department of Biomedical Engineering, Northwestern University, Evanston, IL, USA
b Lawrence Livermore National Laboratory, Medical Technology Program, 7000 East Ave, Livermore, CA, USA
c Lawrence Livermore National Laboratory, Chemistry and Material Science, 7000 East Ave, Livermore, CA, USA
d Department of Bioengineering and Neuroscience, Syracuse University, Syracuse, NY 13244-5290, USA

Received 11 April 2002;  accepted 17 October 2002. ; Available online 5 February 2003.

  1. Abstract

Surfaces of commercially pure titanium and titanium, 6-aluminum, 4-vanadium were subjected to simultaneous polarization/impedance testing and in situ electrochemical atomic force microscopy imaging to evaluate how the structure and properties of the passive oxide film is affected by varying potential and hydration. Current transients were acquired via a step polarization impedance spectroscopy technique: the voltage was stepped between −1 and 1 V in 50 mV increments, while current transients and surface morphology were digitally recorded. Numerical Laplace transformation applied to the current transient data provided frequency-dependent admittance (impedance−1). Simultaneous AFM imaging of dry surfaces, initially hydrated surfaces, and surfaces immersed and changing with potential revealed that all sample surfaces were covered with protective titanium oxide domes that grew in area and coalesced due to hydration and as a function of increasing applied voltage and time. Reversal of dome growth did not occur upon voltage reduction, while impedance behavior was quasi-reversible, suggesting independence between structural and electrical properties. Oxide growth appeared to occur in part by lateral spreading and overgrowth of domes at the oxide-solution interface. Interfacial impedance data reflect oxide passivity and n-type semiconductor behavior. Non-linear Mott-Schottky fits specified multi-layer donor concentrations between 1018 and 1019 cm−3, depending on the surface.

Author Keywords: Titanium; Titanium oxide; Atomic force microscopy; Electrochemical methods; Corrosion; Surface structure; Interfaces
0x01 graphic

  1. Article Outline

1. Introduction

2. Materials and methods

3. Results

4. Discussion

5. Conclusion

Acknowledgements

References


0x01 graphic

  1. 1. Introduction

A unified theory of titanium and its protective oxide, accounting for both its steady state composition and transient behavior, remains elusive. Most previous studies have collected electrochemical data through in situ techniques, while accompanying structural data typically has been collected through ex situ, high vacuum techniques, in which the oxide is no longer in its native environment. Thus, titanium's response—be it morphologically, structurally, electrically, or corrosion resistivity—to transient electric fields and potential shifts in solution is not well understood. A more complete understanding of this response is critical in a variety of applications, including biomedical applications, which are the focus of this study.

By weight or volume, orthopedic implants far exceed any other biomaterial application. These implants often contain titanium and may be used as fracture fixation devices or as joint replacement devices or prostheses. At present, average life span of an artificial hip is approximately 10 years. Total hip replacement procedures are becoming more common in younger patients; patients are therefore exposed to greater mechanical stress over a longer period of time. Increasing the life span of artificial joints is clearly a challenging focus of biomaterial research. Specifically, identifying how to achieve fast, strong, permanent bone-implant fixation is a goal of the orthopedic community. Our focus is on examining nanometer scale implant material characteristics and performance in physiological electrochemical environments related to fixation.

It is the metallic oxide, as opposed to substrate metal, that comes into direct contact with biological milieu, and chemical properties of the oxide may influence interactions with biomolecules. Biocompatibility considerations make characterization of metallic biomaterial surface oxides of utmost importance. This study sought to concurrently obtain data on titanium's electrical properties and surface morphology to understand the relationship between oxide structure and properties and to assess the role of solution exposure and potential variations. Titanium samples were immersed in solution and subjected to step polarization impedance spectroscopy (SPIS) [1] experimentation along with concurrent electrochemical atomic force microscopy (EC AFM) imaging, to more fully characterize titanium's response to transient electric fields and potential shifts.

Oxide structure of commercially pure titanium (CPTi) and titanium, 6-aluminum, 4-vanadium (Ti-6Al-4V) differ [2, 3, 4, 5 and 6]. Previous studies have shown that immersion time, chemical pre-treatment, and potential may affect oxide film structure [7, 8, 9 and 10]. Crystallization may occur on oxides thickened over time, and ion incorporation, including calcium and phosphate influx [7] may take place [7, 11 and 12]. Additionally, hydration causes documentable shifts in binding energies, which may be measured with XPS or Auger [11 and 13]. Changing environmental conditions may also result in open circuit potential (OCP) drops approaching 1 V, with subsequent oxide fracture [14 and 15].

A drawback of all these studies is that they collected surface structure information using high vacuum techniques, thus presenting a very different environment to surfaces than observed in solution. Only electrochemical information was obtained in situ. The differences between hydrated oxide and an oxide in vacuum remain unexamined.

The paucity of surface structure probes operable in solution has impeded characterization of titanium's native surface structure. Although photo-electrochemical microscopy has been used to map grain structure and oxide heterogeneities at a Ti/TiO2 interface, as well as to determine properties of electrochemically grown films [16 and 17], illumination with energy above Ti's bandgap is necessary, and surface physics therefore are altered. Conversely, AFM provides near-atomic level imaging of topographical changes in wet surfaces without such interference. While promising, only basic AFM characterization of Ti/TiO2 has been published to date. Browne et al. [13] used AFM to investigate titanium under various surface treatments and with biolayer coverage, as well as to examine growth of oxide domes and dissolution of titanium metal electrodes immersed in oxalic acid [10].

Turning now to electrochemical characterization, such characterization provides invaluable insight into chemical reactions, potential activity ranges, kinetics, and oxide film parameters (e.g., resistance, thickness). For titanium, relevant electrochemical characterization probes are typically separated into DC polarization-type and AC impedance-type. DC polarization probes provide a first-order determinant of corrosion resistance, while AC electrical impedance spectroscopy (EIS) probes characterize electrochemical interfaces. Pan et al. [18] studied TiO2 with both probe types. They separately performed EIS and polarization tests in phosphate buffered saline (PBS) on titanium and stainless steel substrates. Starting topography of one substrate was characterized with an AFM image, but the morphological changes in the surface were not tracked over the course of the study.

Recent work by the authors of the present study examined the effects of voltage on titanium hydration [19], which prompted inquiry into the relationship between structural and electrical properties of hydrated titanium surfaces. Thus, experimentation was conducted on titanium oxide from CPTi and Ti-6Al-4V. The samples were immersed in PBS and imaged with in situ EC AFM during simultaneous impedance testing. Impedance measurements, collected using SPIS, provided concurrent, inexpensive determination of polarization and electrochemical impedance behavior at an electrochemical interface as a function of potential.

  1. 2. Materials and methods

An experimental method based on slowly stepping potential and concurrently imaging the sample surface while collecting polarization and impedance data was employed to simultaneously characterize the morphology and electrical characteristics of titanium and its oxide film.

Sample preparation, as well as the experimental testing configuration, have been described in a previous publication [19]. Briefly, in situ testing was performed on an electropolished and Kroll's solution etched titanium substrate of CPTi and Ti-6Al-4V. The experimental setup included a glass fluid cell (Digital Instruments), which removably held the silicon nitride cantilever tip (Nanoprobe SPM tip, Digital Instruments). The fluid cell was clamped on top of a titanium test substrate and sealed with an O-ring ( Fig. 1 and Fig. 2). The setup included a three-electrode electrochemical testing configuration, in which the titanium sample acted as a working electrode (WE), a silver wire (99.99% purity, 0.25 mm diameter) provided a quasi-reference electrode (RE), and a platinum wire (99.99% purity, 0.25 mm diameter) served as a counter electrode (CE). The silver wire was inserted through the top of the cell, and the platinum wire was contained in immersed outport tubing. The sample chamber held 50 0x01 graphic
l of solution and was maintained at 25°C.

0x01 graphic

(7K)

Fig. 1. Schematic illustration of underside of electrochemical fluid cell (top view).

0x01 graphic


(12K)

Fig. 2. Side view of EC AFM system design.

0x01 graphic

Using Digital Instruments software (v4.42 r4), surfaces were first imaged dry in contact mode, on the electrochemically etched titanium surface. The images were centered on a distinct surface feature, such as a triple point of grain boundaries. Both height and deflection (the error signal) images were obtained. After disengaging and retracting the tip, PBS was injected into the fluid cell input port via a syringe coupled by luer lock to the port, immersing 0.5 cm2 of titanium, the imaging tip, the reference electrode, and the CE in solution. A potentiostat (Model 362, EG&G Princeton Applied Research, Princeton, NJ) applied −1 V to the titanium surface from the instant of fluid injection. The syringe was left in place during experimentation, and exit port tubing was clamped off to provide backpressure.

The AFM tip was then re-engaged on the surface, and the exact location of dry imaging was reacquired. Si3N4 tips with 200 0x01 graphic
m cantilevers (nominal stiffness of 0.06 N/m), exhibiting an approximate change of 1 V between undeflected and contacted cantilever positions (a positional change of approximately 70 nm/V), exerted a nominal force in the range of 4 nN. This is an approximation; however, higher forces did not influence surface morphology, but lower forces resulted in poor contact and "double-layer" effects. Because the direction of the reflection AFM beam changes slightly when switching from an air to liquid environment, scanning initially resumed over a 50×50 0x01 graphic
m2 area, thus allowing easy identification of the previously analyzed triple point of grain boundaries. Once this area was re-identified, the scanning area was reduced to the original area of interest (typically 5×5 0x01 graphic
m2). Surfaces were held at −1 V for approximately 1 h before SPIS testing commenced [1].

Small steps in potential (50 mV) were then applied to the electrochemical interface with concurrent AFM imaging, and the resulting current transient response was logarithmically collected by a computer from 50 0x01 graphic
s to 10 s, using A-D methods (National Instruments A/D Board, AT-MIO16; and Lab Windows Software, Austin, TX) (Fig. 3 and Fig. 4). The impedance of the interface was determined using a transfer function analysis (the Laplace transform of the current output divided by the Laplace transform of the voltage input) using a previously developed numerical Laplace transform algorithm [1] to determine the frequency-dependent admittance (inverse impedance). By comparing the admittance to a model Randle's circuit ( Fig. 5), minimum and maximum currents per step, high frequency (or early time) resistance (Re), polarization resistance (Rp), capacitance (C), and admittance (A′ and A″) were determined. Current-voltage relationships yielded polarization behavior, charge transfer, and interfacial capacitance.

0x01 graphic

(6K)

Fig. 3. Schematic of experimental EC AFM/SPIS setup where the voltage is stepped in 50 mV increments over the course of approximately 60 min. Current transients as well as an image of the surface are collected at each step.

0x01 graphic


(6K)

Fig. 4. Exemplary current transient collected logarithmically over 10 s resulting from a 50 mV step in voltage from −1 V to—950 mV versus Ag wire on CPTi surface.

0x01 graphic


(2K)

Fig. 5. Schematic of Randle's circuit, where Re represents an oxide resistance, Rp represents polarization resistance, and CMS represents Mott-Schottky capacitance.

0x01 graphic

Capacitance determined from current transient data was used for Mott-Schottky plots of C−2 versus potential to obtain flat band potentials and donor density concentrations. The slope of six or seven initial anodic points on each plot was calculated and used to determine an initial donor density value, N(V), for the non-linear Mott-Schottky fits. Initial N(V) and associated applied voltage and capacitance values were then used to determine N(V) versus voltage for the approximate range of 0-1 V for the CPTi and Ti-6Al-4V.

Though researchers have reported uniform donor concentrations based on linear Mott-Schottky plots [18 and 20], these concentrations only apply to single crystal materials. It is expected that polycrystalline and amorphous TiO2 Mott-Schottky plots will deviate from a linear relationship because of the multiple donor levels which result from lattice defects, grain boundaries, dislocations, etc. in oxide films [21 and 22]. Dean and Stimming [23], Pyun and Kim [22], and Lee and Pyun [24] therefore non-linearly modeled Mott-Schottky by assuming consecutive ionization of donor levels and by utilizing the Euler method [25].

All statistical calculations were performed using Statistica (Statsoft, Tulsa, OK) unless otherwise noted.

  1. 3. Results

Fig. 6 presents two AFM deflection images (10 0x01 graphic
m×10 0x01 graphic
m) of titanium surfaces taken in air after electropolishing and etching. Fig. 6(a) depicts CPTi, while Fig. 6(b) is of Ti-6Al-4V. The CPTi surface of Fig. 6(a) has a hexagonal close packed (HCP) microstructure and is available in four grades of increasing oxygen content. The CPTi surface has three oxide-dome-covered 0x01 graphic
phase grains separated by grain boundaries. Note that the domes appear qualitatively different on the different metal grains. The bottom grain is likely a different crystal plane.

0x01 graphic

(33K)

Fig. 6. AFM deflection images (10 0x01 graphic
m×10 0x01 graphic
m) of titanium surfaces; (a) CPTi and (b) Ti-6Al-4V. Both are electropolished and Kroll etched. The CPTi (a) shows a triple point of grain boundaries and 3 oxide-dome-covered 0x01 graphic
grains. The Ti-6Al-4V (b) contains 0x01 graphic
grains covered with oxide domes and is occasionally interrupted by 0x01 graphic
grains sitting higher on the surface. The oxide morphology appears to be grain dependent (affected by substrate crystallographic orientation).

0x01 graphic

Referring now to Fig. 6(b), Ti-6Al-4V is a bi-modal alloy consisting of globular 0x01 graphic
phase grains surrounding transformed 0x01 graphic
phase. The transformed 0x01 graphic
is comprised of retained 0x01 graphic
(body centered cubic, BCC) and acicular 0x01 graphic
. On the Ti-6Al-4V surface, thermo-mechanical processing disrupts acicular 0x01 graphic
, which results in a high fatigue strength, bi-modal microstructure. Distinct transformed 0x01 graphic
grains interrupt the 0x01 graphic
phase across the Ti-6Al-4V. Needle-like acicular 0x01 graphic
grains in the vicinity of the 0x01 graphic
grains are also apparent. Here too, the morphology of the oxide domes appears to be 0x01 graphic
grain specific (compare lower left to lower right grains), where the height and density of domes varies from grain to grain. Also, the blocky 0x01 graphic
grains do not have oxide domes, but appear more uniform.

The 5 0x01 graphic
m×5 0x01 graphic
m images of Figs. 7(a)-(f) depict transient oxide dome evolution on CPTi as a function of hydration and voltage. Oxide dome broadening, nucleation, and roughening may be imaged, but rising or lowering of the whole surface with respect to a lower plane may not. Therefore, the height of a prominent oxide dome with respect to background height has been tracked as potential was changed (see Table 1). A triple point of grain boundaries on the CPTi surface provides a common reference throughout the voltage sequence. The black arrow points to the specific oxide dome tracked, the height of which is recorded at each voltage condition and listed in Table 1. Image roughness values are also listed. Fig. 7(a) shows the dry surface, where oxide domes on the far left grain are more dispersed than on the other two grains. Fig. 7(b) shows the surface just after immersion in PBS at −1 V, and Fig. 7(c) illustrates the surface after nearly an hour at −1 V. Domes on the far left grain nucleate to shield the entire substrate. Figs. 7(b) and (c) show a progressive swelling of oxide domes with time. Although a cathodic voltage reduces oxide dome growth kinetics as will be shown below, thermodynamics still force the surface to oxidize via hydration, as evidenced by oxide dome growth, even at −1 V. Oxidation via hydration, under both freely corroding and potentiostatically held conditions, is described in a previous paper [19].

0x01 graphic

(28K)

Fig. 7. Sequential AFM deflection images (5 0x01 graphic
m×5 0x01 graphic
m) of electropolished and Kroll etched CP titanium. All images include the same triple point of grain boundaries; (a) dry, (b) after initial hydration in PBS at −1 V, (c) held at −1 V for ~1 h showing nominal oxide dome growth, (d) after at 0 V, (e) titanium at +1 V, and (f) titanium returned to −1 V. Note that increasing in potential leads to drastic oxide dome growth and coalescence. Cathodic ramping of voltage back to −1 V does not cause oxide domes to shrink. The arrow indicates oxide dome used for specific height measurement (Table 1).

0x01 graphic

Table 1. Relative oxide dome height of CPTi feature (see arrow in Fig. 5) recorded as a function of potential in PBS, and entire image roughness as a function of applied potential (5 0x01 graphic
m×5 0x01 graphic
m)
0x01 graphic

0x01 graphic

Subsequent slow, linear ramping of potential to 1 V results in observable, increased lateral oxide dome growth, as well as coarsening, across the grains and grain boundaries, as seen in Figs. 7(d) and (e). The oxide dome area of the initially immersed surface at −1 V is 1.4×104 nm2±1.1×104 nm2; the −1 V oxide dome area at 1 h is 1.8×104 nm2±1.2×104 nm2, and the +1 V area is 1.8×105 nm2±7.5×104 nm2 0x01 graphic
m2 (n=30 domes). Dome area increased approximately 1.47 times at −1 V (~1 h), and approximately 5.31 times between the −1 and +1 V.

The −1 V area data is statistically different than the +1 V data (p=0.0001 Anova, post hoc Newman-Keuls). The two −1 V data sets, meanwhile, are more similar (p=0.76 Anova, post hoc Newman-Keuls). Clearly, dome broadening cannot be eliminated through application of potential within the described regime, but initial hydration conditions influence it. The word initial is stressed, because reversion to an applied potential of −1 V over approximately 30 min, as seen in Fig. 7(f), did not cause dome area to decrease. Films are thought to thin under cathodic bias, but AFM is unable to yield absolute height data in these experiments. Irreversible expansion of dome area may seem surprising and is a focal point of the discussion section.

As mentioned above, Table 1 includes relative height changes of a specified oxide dome with respect to background (see arrow series in Fig. 7 for identification of dome tracked with potential). The dome was on a grain boundary and was chosen because it was the easiest dome to follow through the entire series of images. Vertical distance between dome peak height and background floor first increased upon immersion in solution, presumably due to the oxide dome growing as the floor remained constant. As oxidation via hydration increased, however, the relative height of the dome decreased with time and increasing potential. It appears that as oxide domes nucleate to cover the entire surface, then impinge and coalesce, the relative oxide floor rises, the boundaries between oxide domes decrease in number and in height, and relative heights therefore decrease. Reversion of the potential to −1 V did not significantly alter height in the particular dome tracked.

Roughness first decreased upon immersion in solution, probably due to dome impingement and initial coalescence, which served to lessen the height variation between all domes. Maintaining a cathodic potential on the surface of the sample for an extended period of time slightly increased roughness, and raising the potential slightly decreased roughness, but the absolute changes are so small (subnanometer) that they are not thought to be significant.

Oxide dome areas as a function of potential and time were quadratically fitted (Figs. 8(a) and (b)). The R2 value for area versus potential is 0. 89 and the R2value for area versus time is 0.90. The quadratic relationships are assumed based on the fact that potential was increased during the experiments (increasing growth rate with increasing potential). Had potential been held constant, a linear relationship would have been expected.

0x01 graphic

(16K)

Fig. 8. (a) Average oxide dome area as a function of potential (R2=0.90); (b) average oxide dome area as a function of time (R2=0.89). (Quadratic fit according to Y=Y0+aX+bX2; for (a) Y0=7.529×104, a=−1.727×103, b=1.486×101; for (b) Y0=4.94×104, a=6.108×104, b=4.026×104.)

0x01 graphic

Referring back to the Ti-6Al-4V sample, as seen in the sequential images of Fig. 9, the Ti-6Al-4V surface responded to hydration and applied voltage in a manner similar to CPTi. The sequence begins with the dry 5 0x01 graphic
m×5 0x01 graphic
m AFM image of Fig. 9(a) and then shows the same region immersed in PBS at −1 V in Fig. 9(b). Oxide domes on the 0x01 graphic
and acicular 0x01 graphic
phases nucleate and impinge on the surface upon hydration. Hydration does not initially lead to a notable increase in dome area. However, by the time the voltage was increased to -0.5 V (Fig. 9(c)), the flat substrate was no longer visible, and all areas were protected with domes.

0x01 graphic

(31K)

Fig. 9. AFM deflection images (5 0x01 graphic
m×5 0x01 graphic
m) of electropolished and Kroll-etched Ti-6Al-4V. All images include identical central region of needle-like acicular 0x01 graphic
phase extending from a 0x01 graphic
grain on a predominant 0x01 graphic
phase background. Downward drift in images is noted; (a) dry titanium, (b) titanium initially hydrated in PBS at −1 V, (c) titanium at -0.5 V, (d) titanium at 0 V, (e) titanium at +1 V, and (f) titanium returned to −1 V. Again, increase in potential leads to drastic oxide dome growth and coalescence. Oxide dome growth is prevalent on 0x01 graphic
phase; growth of 0x01 graphic
features is also seen. Cathodic ramping of voltage back to −1 V decreases height but not lateral dimensions on 0x01 graphic
features. See text for arrow explanation.

0x01 graphic

0x01 graphic
phase change was also visible. Area of 0x01 graphic
grains increased, but oxide domes were not apparent. Rather, the surface oxide was smooth on the top and ridged on the sides (Fig. 9). When potential was anodically ramped, the oxide domes expanded on the 0x01 graphic
and acicular 0x01 graphic
phases. 0x01 graphic
grain oxides grew upon anodic ramping as well, and oxide covering these grains appeared to grow out over the adjacent 0x01 graphic
phase. However, it is unknown whether 0x01 graphic
oxide predominantly covers 0x01 graphic
oxide, or if the two phases are miscible. Fig. 9 illustrates the progression as voltage was ramped to -0.5 V (Fig. 9(c)), 0 V (Fig. 9(d)), 1 V (Fig. 9(e)), and finally back to −1 V (Fig. 9(f)). As with the CPTi surface, reducing the voltage to −1 V did not result in surface structure reversion (Fig. 9(f)). Although downward drift is apparent in the image sequence, at least three 0x01 graphic
grains may be followed. A large expansion in 0x01 graphic
grain area is clearly visible.

Similarly to Fig. 7, Fig. 9 contains arrows pointing to both an 0x01 graphic
phase oxide dome and a 0x01 graphic
grain in each image of the potential-dependent series presented in Fig. 9. Relative heights of these features, as well as image roughness values, are listed in Table 2. Some tip convolution is apparent on the anodized images, but post-experimental analysis confirmed that the AFM remained calibrated and that the tip maintained sufficient resolving power. It should also be noted that as domes spread and flatten, convolution should decrease. Thus, more convolution would be expected at low applied voltages than at higher applied voltages and later times.

0x01 graphic

Table 2. Relative heights of 0x01 graphic
phase oxide dome and 0x01 graphic
phase grain of Ti-6Al-4V features (see arrows in Fig. 9) recorded as a function of potential in PBS, and entire image roughness as a function of applied potential (5 0x01 graphic
m×5 0x01 graphic
m)
0x01 graphic

0x01 graphic

Due to the inclusion of the 0x01 graphic
phase, these results differ from the CPTi results. The isolated 0x01 graphic
phase oxide dome examined had a comparable relative height to the one measured in Fig. 7, but anodic ramping caused such drastic coalescence that it became difficult to decipher oxide dome boundaries by +1 V. Accordingly, the height of the 0x01 graphic
phase oxide dome on the Ti-6Al-4V surface at 1 V, as listed in Table 2, reduced to 2.3 nm. In contrast, the 0x01 graphic
phase oxide dome examined on the CPTi surface at 1 V was still 18.4 nm tall. However, examined grains were chosen according to which ones could be tracked across the entire voltage series with confidence. The 0x01 graphic
phase oxide is thought to behave similarly on the two substrates and Table 1 and Table 2 both indicate a trend of coalescing oxide leading to a decrease in relative height between domes.

The 0x01 graphic
phase grain selected grew considerably in its lateral dimensions, from 478.5 to 830.1 nm (173.5%) along the top edge, and increased ~17 nm (117.4%) in the z-direction. Height increased slowly upon substrate immersion and anodic ramping of potential. The largest height change was seen when potential increased from 0 to +1 V. Notably, cathodic ramping back down to −1 V decreased the 0x01 graphic
phase grain height by 9 nm. This suggests increased reactivity of the 0x01 graphic
phase, as compared to the 0x01 graphic
phase, within the sample. Roughness of the Ti-6Al-4V sample is dominated by the tall 0x01 graphic
phase grains and is therefore greater than that of CPTi. Roughness first decreased upon immersion in solution, but then kept increasing with anodic ramping. This suggests a trend of roughness changing directly with 0x01 graphic
grain height during anodic ramping.

With reference now to Fig. 10, Fig. 11, Fig. 12, Fig. 13, Fig. 14, Fig. 15 and Fig. 16, electrical data is presented. Fig. 10 and Fig. 11 show the in-phase admittance, A′, of the CPTi and Ti-6Al-4V samples, respectively, as a function of frequency and voltage, while Fig. 10 and Fig. 11 illustrate out-of-phase admittance, A″, of the CPTi and Ti-6Al-4V samples, respectively, as a function of frequency and voltage within the EC AFM fluid cell. For both surfaces, in-phase admittance peaks at highest frequency before decaying with decreasing frequency and increasing voltage. Peak values are: 8064 0x01 graphic
S/cm2 for CPTi (4136 rad/s), and 14196 0x01 graphic
S/cm2 for Ti-6Al-4V (6990 rad/s). With respect to voltage, CPTi rolls off gradually, and Ti-6Al-4V drops to approximately 7000 0x01 graphic
S/cm2 and then rolls off. These results are similar to those obtained for CPTi in a previous study, but slightly lower in amplitude, suggesting a thicker oxide film [1].

0x01 graphic

(31K)

Fig. 10. Three-dimensional plot of (a) in-phase (A′) and (b) out-of-phase (A″) admittance of CPTi in PBS versus frequency (rad/s) and potential (V).

0x01 graphic


(30K)

Fig. 11. Three-dimensional plot of (a) in-phase (A′)and (b) out-of-phase (A″) admittance of Ti-6Al-4V in PBS versus frequency (rad/s) and potential (V).

0x01 graphic


(20K)

Fig. 12. Polarization plot and impedance characteristics for CPTi in PBS. Polarization plot (a) indicates maximum currents (50 0x01 graphic
s after potential steps) and minimum currents (10 s after potential steps) versus voltage. Interfacial capacitance, C, is plotted in (b) versus voltage. Resistance in (c) is plotted as early (Re), calculated from polarization data, or high frequency (Rs), calculated from admittance data, versus voltage; resistance in (d) is plotted as late (RL), calculated from polarization data, or polarization (Rp), calculated from admittance data, versus voltage. Capacitance was calculated as, C=(Rs+Rp)/RsRp0x01 graphic
A, where 0x01 graphic
A is the frequency (rad/s), where the peak of the loss admittance, A″ occurs. Early (Re) and solution (Rs) resistance were calculated as, ReV/(ImaxIinitial), and0x01 graphic
. Late (RL) and polarization (Rp) resistances were calculated as: RLV/(IminIinitial), 0x01 graphic
.

0x01 graphic


(20K)

Fig. 13. Polarization plot and impedance characteristics for Ti-6Al-4V in PBS. Polarization plot (a) depicts maximum currents (50 0x01 graphic
s after potential steps) and minimum currents (10 s after potential steps) versus voltage. Interfacial capacitance, C, is plotted in (b) versus voltage. Resistance in (c) is plotted as early Re and Rs versus voltage; resistance in (d) is plotted as RL and Rp versus voltage.

0x01 graphic


(8K)

Fig. 14. Calculated and experimental charge versus voltage step number. Each step was 50 mV; potential was ramped linearly from −1 to +1 V (step 40) and back down to −1 V (step 80). Calculated charge calculated according to: QX0x01 graphic
ZFA/MW, where ΔX=0x01 graphic
(VVi), 0x01 graphic
is the anodization rate for titanium (2 nm/V), and VVi is 50 mV; 0x01 graphic
is density of the oxide (4.4 g/cm3); Z is the valence state of the oxide (+4); F is Faraday's constant (96,480 c/mol); A is electrode area (0.5 cm2); and MW is molecular weight of TiO2 (79.9 g/mol). The area under the current transient curves equals experimental charge and was determined by trapezoidal Reimann sum approximation.

0x01 graphic


(5K)

Fig. 15. Mott-Schottky plot of C−2 versus voltage for CPTi anodically ramped from −1 to +1 V over 40 steps. Note the curve on the right where the applied potential is positive. This is due to multiple donor levels in a polycrystalline sample.

0x01 graphic


(5K)

Fig. 16. Donor concentration (cm−3) as a function of voltage for the CPTi and Ti-6Al-4V surfaces, as calculated by the non-linear Mott-Schottky fit.

0x01 graphic

Out-of-phase admittance curves of Figs. 10-11(b) are bell shaped. On both surfaces, peak values occur at characteristic frequencies and decrease with increasing frequency and voltage (see discussion). Peak A″ values have associated characteristic frequencies of relaxation of 390 rad/s at -0.95 V, 507 rad/s at -0.5 V, 857 rad/s at 0 V, and 2447 rad/s at 1 V for CPTi; and 177 rad/s at -0.95 V, 390 rad/s at -0.5 V, and 659 rad/s at 0 V, and 845 rad/s at 1 V for Ti-6Al-4V.

Fig. 12 and Fig. 13 plot polarization data for maximum currents (50 0x01 graphic
s) and minimum currents (10 s), after each step in potential for the CPTi and Ti-6Al-4V surfaces. These minimum current results are similar to those measured in standard potentiodynamic polarization tests. Minimum current polarization plots all contain a potential where current approaches zero, referred to as zero current potential, as well as a passivity region where current stays constant with increasing voltage (approximately -0.5 and +1 V for CPTi and Ti-6Al-4V). This passivity region is characteristic of valve metals like titanium. Maximum current polarization plots are similar for both surfaces, although more hysteresis is present in the CPTi curve than in the Ti-6Al-4V curve. Maximum currents also allude to the passivity region, but bowing of the curves at roughly -0.5 V is not nearly as defined as the demarcation of Corrosion potential, Ecorr, from minimum current readings.

Early/solution resistances also reflect oxide film passivity, as seen in Figs. 12-13(c). Thus, early resistance (Re) appears to be a more appropriate label than the traditional solution resistance label (Rs). Anodic voltage ramping yielded the lower curves of Figs. 12-13(c), and cathodic ramping yielded the top curves. Re clearly varies with potential, even though resistances continued to increase when potential ramping was first reversed from +1 V toward −1 V, and hysteresis is present on all surfaces. Nevertheless, starting and ending resistance values are the same. Late (RL) and polarization (Rp) resistances are noisy, but also similar (see Fig. 12 and Fig. 13).

Capacitance values started high at approximately 100 0x01 graphic
F/cm2 and decreased with increasing voltage to approximately 1 0x01 graphic
F/cm2. The top curves of Fig. 12 and Fig. 13 correspond to increasing voltage from −1 to +1 V; the bottom curves correspond to voltage ramping back down to −1 V. When voltage was first reversed, capacitance temporarily continued to decrease, and hysteresis is present in all three curves in a similar, yet opposite, trend to the early (solution) resistance plots. In contrast to oxide dome expansion, which is irreversible while surfaces are maintained in a hydrated state, electrical properties are reversible.

Fig. 14 provides calculated and experimental charge for all voltage steps on the Ti-6Al-4V sample. Current transients require charge transfer through the oxide film. Hysteresis present in the current-voltage relationships of Fig. 12 and Fig. 13 initially suggested that current transients were not taken over a sufficient time interval. However, if 10 s had not been long enough, currents transients should not have indicated such drastic decay (as in Fig. 9).

Fig. 14 shows that, on average, the measured charge was one-fifth of the calculated charge needed for oxide growth. Early and late steps, roughly corresponding to steps between −1 and -0.5 V and then between -0.5 and −1 V, indicate higher charge transfer. The time constant, 0x01 graphic
, was calculated as the inverse of 0x01 graphic
A, the peak of admittance, A″, by plotting A″ versus log 0x01 graphic
. Four times 0x01 graphic
describes a time process that is 98.2% complete and five times 0x01 graphic
describes a nearly completed process [26]. Five 0x01 graphic
corresponds to 16 ms in the above experiment. Therefore, current acquisition time was sufficient, even though experimental values were less than calculated.

Finally, Fig. 15 presents an exemplary non-linear Mott-Schottky plot of initial C−2 versus voltage for CPTi, while Fig. 16 plots N(V) versus V for both surfaces. Dielectric constant for TiO2 was held at 50. Equations used are valid for all potentials positive of the flatband potentials. Thus, values between 0 and 1 V are plotted, as the flatband potentials were -0.01 V for CPTi, and 0.073 V for Ti-6Al-4V. A range of flatband potentials for TiO2 have been reported in the literature; typically, values between −0.3 V [20 and 27] and −0.5 V [18] are reported. The non-linearity of polycrystalline samples makes it difficult to accurately assign flatband potentials to Mott-Schottky plot data. It is thought that the consistent strategy of taking the first six anodic points from the C−2 versus potential plots for initial N(V) values, may have slightly increased flatband potential calculations. However, using fewer data points would have introduced more error into initial N(V) values. CPTi donor values were on the order of 1018, and Ti-6Al-4V values were on the order of 1019.

  1. 4. Discussion

Simultaneously probing structural and electrical characteristics of titanium oxide films allowed direct observation of lateral dome spreading, the voltage and time dependence of oxide processes, as well as notable surface roughness changes. Specifically, the oxide growth rates discussed previously reveal a quadratic relationship between dome area and increasing, positive applied potential.

Our previous examination of Ti/TiO2 hydration under various electrochemical conditions [19] revealed that applying a static cathodic potential results in an oxide area growth rate that is slower than the corresponding rate at quasi-static potential, near 0 V OCP. In this study, ramping of applied potential in the anodic direction results in a constantly changing growth rate. However, assuming a linear relationship in the region of applied potential around ~0 V, a rate of 1070 nm2/min is obtained (R2=0.77); this value closely approximates the OCP rate obtained previously (1098 nm2/min) [19].

Data also reveal reversible electrical responses with respect to voltage, but irreversible structural response. Structurally, dome area continually increases with time and voltage for all surfaces, but does not decrease over time with decreased voltage. This independence of structure and electrical properties after hydration was initially unexpected. However, as qualitatively modeled in previous TiO2 hydration experiments [19], TiO2 is classified as an n-type semiconductor, and its oxide primarily grows and changes at the metal/oxide interface, not the oxide/solution interface, where image acquisition occurred. While new oxide is forming, existing oxide domes at the metal/oxide interface expand laterally, and material is pushed up from below. What was once a surface with many small domes, changes into a surface exhibiting only a few broad domes. Upon ramping potential cathodically, the broad oxide domes at the oxide/solution interface, located nanometers above the metal/oxide interface, do not transform back to the multiple small domes. Rather, the oxide domes just appear to flatten slightly. Much more significant remodeling is anticipated at the metal/oxide interface. However, since EC AFM probes morphology at the oxide/solution interface, electrical changes are more apparent than structural changes during cathodic ramping. Future work X-ray experiments are required to probe morphological changes at the metal/oxide interface.

The one noted exception to the surface achieving steady state morphology independent of applied, or more specifically reduced, potential is the 0x01 graphic
phase of Ti-6Al-4V surfaces. 0x01 graphic
phase grain heights and applied potential are proportional, as provided in Table 2. This should be considered with respect to vanadium dissolution observed and reported in physiological systems [28]; high vanadium concentrations in persons with Ti-6Al-4V orthopedic implants may result from rapid growth and dissolution of 0x01 graphic
grain oxide. Comparing changes in 0x01 graphic
phase grain heights to the negligible height changes noted in 0x01 graphic
phase oxides during ramping of applied potential from anodic to cathodic conditions (where Al is more soluble) supports this.

The decoupled nature of surface morphology and electrical properties may be useful in biological systems, due to the presence of a more stable interface for an adsorbed biolayer. TiO2 is modeled as initiating growth or recession/reduction at the metal/oxide interface. However, the solution/oxide interface initially presents growing oxide domes to the physiological environment. It is therefore hypothesized that osseointegration begins during lateral oxide dome growth, incorporating selected material into the oxide.

As Zitter and Plenk described [29], electron flow in orthopedic implant metal systems must equal ion flow through tissue and biological fluids. As an electric field arises assisting oxide growth at the expense of the substrate metal and transported electrons ionize oxygen atoms, adsorption of H2O, OH, HPO42−, and H2PO4− may occur at the oxide/solution interface [19 and 30]. Additionally, the protein albumin is known to bind to titanium in significant amounts [31, 32 and 33], and is the major calcium binding protein in the blood [31]. The Ca2+ ions bridge the negatively charged albumin and the TiO2 surface [31]. Such calcium and phosphate ion incorporation has already been suggested to contribute to oxide growth [34] and provide a gradual transition from metal to living cells [35]. However, this has not been documented on a nanometer scale, in which titanium oxide domes and biological moieties such as proteins interact. As well, ion, protein, and oxide integration may be relatively impervious to mild load related cathodic conditions placed on titanium implants. Even if cathodic potentials are applied to the oxide, interlocking bonds incorporating the material are expected to maintain integrity; instead, bonds at the metal/oxide interface would be expected to restructure. Obviously, integrity might be compromised under conditions of wear, which could result in direct removal of oxide, and thus interlocked regions, starting from the surface and moving inward.

Electrical properties follow expected trends for oxide growth, although hysteresis is present and charge transfer is less than calculated. Early resistance, thought to correspond to a combination of oxide resistance, which increases with potential, and solution resistance, which is constant, is directly proportional to potential, which means that oxide film resistance should increase and decrease with voltage, due to oxide growth and dissolution. Early resistance steadily increases upon anodic ramping (suggesting that oxide dome lateral growth is directly proportional to early resistance), but exhibits a kink upon cathodic ramping, where two different slopes may comprise the voltage reduction curve (Fig. 12).

Due to aspects of the reductive dissolution process that are not fully understood, film growth apparently is not fully reversible, as evidenced by hysteresis. Some oxide growth may occur via the reaction

Ti+O2→TiO2.

(1)

No electrons are transferred in this reaction, but a thermodynamic drive to spontaneously form oxide exists (associated free energy: ΔG=−82.92 kcal/mol). The change in slope upon cathodic ramping and charge increase between -0.5 and −1 V can be attributed to reduction of titanium in a Ti4+ valence state to a Ti3+ valence state. Lower experimental charge at potentials above -0.5 V suggests that non-Faradaic events (e.g., double-layer charging) play an important role, and that oxide thickening between 0 and +1 V acts to block charge transfer. According to Kelly [36], applied anodic potential may oxidize titanium to Ti3+ ions, in addition to Ti4+ions, and cathodic potential may cause reduction of Ti4+ ions to Ti3+ ions. Furthermore, Ti3+ ions are not present at voltages greater than 0 V and become increasingly soluble with application of cathodic potential. Thus, TiO2 films may thin during cathodic potential ramping as reductive dissolution of Ti4+ to Ti3+ takes place. Some researchers have reported this transition as being responsible for an observed blue coloration [37, 38 and 39]. Since experiments were conducted in neutral pH solutions, onset potentials for reduction reactions were greatly reduced. However, even at cathodic potentials, substrate metal never contacts the solution directly because lower order films, including TiH2 and Ti2O3, are present even at active potentials [29]:

2TiO2,hydr+2H++2e0x01 graphic
Ti2O3,hydr+2H2O,Eo=−0.091 V,

(2)

TiO2+6H++6e0x01 graphic
TiH2+2H2O,Eo=−0.47 V.

(3)

Considering all samples were subjected to a cathodic overpotential for close to an hour before SPIS experiments commenced, it is likely that these hydride and lower order oxide films coexist with TiO2.

Dolata et al. [40] studied rutile and anatase TiO2 film impedance. They noted the high electron mobility of anatase films, and that conduction in doped anatase films undergoes metallic transition. Donor concentrations in anatase films were on the order of 5×1018 cm−3. Their theoretical model anatase data coincides with the experimental donor concentration data of Fig. 16. When examined in conjunction with the data of Blackwood [27] suggesting an intermediate layer, as well as conclusions from the previously mentioned Pan article [18], this information further supports that CPTi's oxide is multi-layered, with an anatase-like uppermost TiO2 layer. We believe that the intermediate layer is Ti2O3. This Ti/Ti2O3/TiO2 physical model assumes that Ti2O3 has a lower resistance than TiO2.

SPIS also provided admittance and capacitance data, which may be related to oxide film growth. A′ values consistently decrease with applied potential, consistent with oxide growth. Peak A″ values have associated 0x01 graphic
A values, their characteristic frequencies of relaxation. A″ data is inversely proportional to 0x01 graphic
A. Large 0x01 graphic
A values imply nominal current passage through an oxide film, which in turn implies a thick oxide film; conversely, as film thickness increases, film resistance increases, and less current is passed. The SPIS method, and associated Laplace transforms, assumed a Randle's circuit consisting of an oxide resistor, Re, in series with a parallel connection of a double-layer capacitor and a polarization resistor, Rp (Fig. 4). Capacitance can thus be modeled as

0x01 graphic

(4)

Clearly, as 0x01 graphic
A increases, capacitance decreases, again associated with thickening of the oxide film. We note that capacitance data was used to determine donor density data. For the sake of simplicity, 0x01 graphic
, dielectric constant, was held at 50 for these calculations. However, we hypothesize that 0x01 graphic
is actually a function of applied potential in the solution-oxide-metal construct.

The reported early resistance and current data is reminiscent of magnetic or ferroelectric hysteresis plots of polarization versus electric field. A model analogous to piezoelectricity is therefore postulated for oxide electrical response. Oxidation at cathodic potentials yields some Ti3+ ions, in addition to the more abundant Ti4+. Most likely, as experiments proceed from −1 to 1 V, Ti3+ manifests itself as the Ti2O3 intermediate layer, while Ti4+ manifests as TiO2. Subsequent cathodic ramping from 1 V does not physically or electrically change the oxide until cathodic potentials sufficient to reduce Ti4+ to Ti3+ are reached. Then, some of the TiO2 is reduced to Ti2O3, and oxide resistance decreases. In a piezoelectric, electrical potential controls dimension and strain; in an oxide, electrical potential may control interfacial oxide state. In fact, strain may also be controlled via electrical potential in the oxide, as growth of the interfacial oxide Ti2O3 layer is also expected to reduce strain. Blackwood [27] purported an intermediate thin film that formed irreversibly. Such irreversible growth could potentially act as a micromechanical lock. As mentioned previously, future work X-ray experiments are necessary to analyze oxide film layer composition as a function of depth, although required sensitivity of a few atomic layers may prove challenging.

  1. 5. Conclusion

For the first time, polarization, impedance, and surface structure behavior are concomitantly reported after simultaneously monitoring pure and alloy titanium electrode samples. Lateral oxide dome spreading was directly observed, and changes in surface roughness were noted. Early resistance and interfacial capacitance data show interesting hysteresis and reflect passive oxide characteristics, as well as trends toward growth and dissolution. Initial oxide growth processes depend on time and voltage. Although initial hydration events may be controlled with applied potential, subsequent hydrated titanium surface structure acts independently of electrical properties. This phenomenon may be due to titanium oxide's n-type semiconductor classification; growth and dissolution apparently occur at the metal/oxide interface, as opposed to the oxide/solution interface.
0x01 graphic

  1. Acknowledgements

This work was performed under the auspices of the US Department of Energy by Lawrence Livermore National Laboratory (LLNL) under contract W-7405-Eng.48. The Medical Technology Program and the Materials Research Institute at LLNL provided funding. We graciously thank Robert Haskins at DePuy for the commercially pure titanium and the titanium, 6-aluminum, 4-vanadium.
0x01 graphic

  1. References

1. J.L. Gilbert , Step-polarization impedance spectroscopy of implant alloys in physiologic solutions. J Biomed Mater Res 40 (1998), pp. 233-243. Abstract-MEDLINE | Abstract-EMBASE   | Full Text via CrossRef

2. Lausmaa J, Mattsson L, Rolander U, Kasemo B. Chemical composition and morphology of titanium surface oxides. Material Research Society Symposium, Boston, MA, 1986. p. 351-9.

3. M. Ask, J. Lausmaa and B. Kasemo , Preparation and surface spectroscopic characterization of oxide films on Ti6Al4V. Appl Surf Sci 35 (1988), pp. 283-301.

4. M. Ask, J. Rolander, J. Lausmaa and B. Kasemo , Microstructure and morphology of surface oxide films on Ti-6Al-4V. J Mater Res 5 8 (1990), pp. 1662-1667. Abstract-INSPEC | Abstract-Compendex  

5. J. Lausmaa, B. Kasemo, U. Rolander, L.M. Bjursten, L.E. Ericson, L. Rosander and P. Thomsen , Preparation, surface spectroscopic and electron microscopic characterization of titanium implant materials. In: B.D. Ratner, Editor, Surface characterization of biomaterials, Elsevier, Amsterdam (1988), pp. 161-174.

6. J. Lausmaa, B. Kasemo, J. Mattsson and H. Odelius , Multi-technique surface characterization of oxide films on electropolished and anodically oxidized titanium. Appl Surf Sci 45 (1990), pp. 189-200. Abstract

7. E.A.B. Effah, P.D. Bianco and P. Ducheyne , Crystal structure of the surface oxide layer on titanium and its changes arising from immersion. J Biomed Mater Res 29 (1995), pp. 73-80. Abstract-MEDLINE | Abstract-Compendex | Abstract-EMBASE  

8. A. Aladjem , Review: anodic oxidation of titanium and its alloys. J Mater Sci 8 (1973), pp. 688-704. Abstract-INSPEC | Abstract-Compendex  

9. A. Wisbey, P.J. Gregson, L.M. Peter and M. Tuke , Effect of surface treatment on the dissolution of titanium-based implant materials. Biomaterials 12 (1992), pp. 470-473.

10. G.M. Brown, T. Thundat, D.P. Allison and R.J. Warmack , Electrochemical and in situ atomic force microscopy and scanning tunneling microscopy investigations of titanium in oxalic acid solution. J Vac Sci Technol A 10 (1992), pp. 3001-3006. Abstract-INSPEC   | Full Text via CrossRef

11. K.E. Healy and P. Ducheyne , Oxidation kinetics of titanium thin films in model physiologic environments. J Colloid Interface Sci 150 (1991), pp. 404-417.

12. J.E. Sundgren, P. Bodo and I. Lundstrom , Auger electron spectroscopic studies of the interface between human tissue and implants of titanium and stainless steel. J Colloid Interface Sci 110 (1986), pp. 9-20. Abstract-Compendex  

13. M. Browne, P.J. Gregson and R.H. West , Characterization of titanium alloy implant surfaces with improved dissolution resistance. J Mater Sci: Mater Med 7 (1996), pp. 323-329. Abstract-EMBASE | Abstract-Compendex  

14. J.L. Gilbert and J.J. Jacobs In: M.B. Mayor, Editor, Modularity of orthopedic implants, ASTM, Philadelphia, PA (1997), pp. 45-59.

15. Goldberg JR, Buckley CA, Jacobs JJ, Gilbert JL. In: Parr JE, Marlow DE, editors. Modularity of orthopedic implants. Philadelphia, PA: ASTM, 1996.

16. M.R. Kozlowski, P.S. Tyler and W.H. Smyrl , Photoelectrochemical microscopy of oxide films on metals: Ti/TiO2 interface. Surf Sci 194 (1988), pp. 505-530. Abstract

17. U. Stimming , Photoelectrochemical studies of passive films. Electrochim Acta 31 (1986), pp. 415-429. Abstract

18. J. Pan, C. Leygraf, D. Thierry and A.M. Ektessabi , Corrosion resistance for biomaterial applications of TiO2 films deposited on titanium and stainless steel by ion-beam sputtering. J Biomed Mater Res 35 (1977), pp. 309-318.

19. J.P. Bearinger, C.A. Orme and J.L. Gilbert , Direct observation of hydration of TiO2 on Ti using AFM: freely corroding versus potentiostatically held. Surf Sci 491 3 (2001), pp. 370-387. SummaryPlus | Full Text + Links | PDF (2172 K)

20. R.M. Torresi, O.R. Camara and C.P. De Pauli , Influence of the hydrogen evolution reaction on the anodic titanium oxide film properties. Electrochim Acta 32 (1987), pp. 1357-1363. Abstract

21. W. Schmickler Interfacial electrochemistry, Oxford University Press, New York (1996).

22. S.I. Pyun and C.H. Kim , Determination of donor concentration in anodically passivating polycrystalline TiO2 films from analysis of nonlinear Mott-Schottky plots. Int J Hydrogen Energy 16 (1991), pp. 661-664. Abstract

23. M.H. Dean and U. Stimming , The electronic properties of disordered passive films. Corrosion Sci 29 (1989), pp. 199-211. Abstract

24. E.J. Lee and S.I. Pyun , Analysis of nonlinear Mott-Schottky plots obtained from anodically passivating amorphous and polycrystalline TiO2 films. J Appl Electrochem 22 (1992), pp. 156-160. Abstract-Compendex  

25. R.W. Daniels An introduction to numerical methods and optimization techniques, North-Holland, New York (1978).

26. T.G. Beckwith, R.D. Marangoni and J.H. Lienhard Mechanical measurements, Addison-Wesley, Reading, MA (1993).

27. D.J. Blackwood, R. Greef and L.M. Peter , An ellipsometric study of the growth and open-circuit dissolution of the anodic oxide film on titanium. Electrochim Acta 34 (1989), pp. 875-880. Abstract

28. J.L. Woodman, J.J. Jacobs, J.O. Galante and R.M. Urban , Metal ion release from titanium-based prosthetic segmental replacements of long bones in baboons: a long-term study. J Orthop Res 1 (1984), pp. 421-430. Abstract-MEDLINE  

29. H. Zitter and H. Plenk, Jr. , The electrochemical behavior of metallic implant materials as an indicator of their biocompatibility. J Biomed Mater Res 21 (1987), pp. 881-896. Abstract-MEDLINE | Abstract-Compendex | Abstract-EMBASE  

30. K.E. Healy and P. Ducheyne , Oxidation kinetics of titanium thin films in model physiologic environments. J Colloid Interface Sci 2 250 (1992), pp. 404-417.

31. A. Klinger, D. Steinberg, D. Kohavi and M.N. Sela , Mechanism of adsorption of human serum albumin to titanium in vitro. J Biomed Mater Res 36 (1996), pp. 387-392.

32. B. Walivaara, B-.O. Aronsson, M. Rodah, J. Lausmaa and P. Tengvall , Titanium with different oxides: in vitro studies of protein adsorption and contact activation. Biomaterials 15 10 (1994), pp. 827-834. Abstract

33. G.C.F. Clark and D.F. Williams , The effects of proteins on metallic corrosion. J Biomed Mater Res 16 (1982), pp. 125-134. Abstract-Compendex | Abstract-EMBASE | Abstract-MEDLINE  

34. J.E. Sundgren, P. Bodo and I. Lundstrom , Auger electron spectroscopic studies of the interface between human tissue and implants of titanium and stainless steel. J Colloid Interface Sci 110 (1986), pp. 9-20. Abstract-Compendex  

35. K.E. Healy and P. Ducheyne , Hydration and preferential molecular adsorption on titanium in vitro. Biomaterials 13 8 (1992), pp. 553-561. Abstract

36. E.J. Kelly In: J.O.M. Bockris, B.E. Conway and R.E. White, Editors, Modern aspects of electrochemistry, Plenum Press, New York (1982), pp. 319-324.

37. C.K. Dyer and J.S.L. Leach , Reversible reactions within anodic oxide films on titanium electrodes. Electrochim Acta 23 (1978), pp. 1387-1394. Abstract

38. R.M. Torresi, O.R. Camara and C.P. De Pauli , Influence of the hydrogen evolution reaction on the anodic titanium oxide film properties. Electrochim Acta 32 (1987), pp. 1357-1363. Abstract

39. J. Pan, D. Thierry and C. Leygraf , Hydrogen peroxide toward enhanced oxide growth on titanium in PBS solution: blue coloration and clinical relevance. J Biomedical Mater Res 30 (1996), pp. 393-402. Abstract-MEDLINE | Abstract-EMBASE | Abstract-Compendex   | Full Text via CrossRef

40. M. Dolata, P. Kedzierzawski and J. Augustynski , Comparative impedance spectroscopy study of rutile and anatase TiO2 film electrodes. Electrochim Acta 41 (1996), pp. 1287-1293. Abstract | Abstract + References | PDF (583 K)
0x01 graphic

Corresponding author. Tel.: +1-315-443-2105; fax: +1-315-443-9175

1 Presently at ETH Zurich, Institut für Biomedizinische Technik und Departement Werkstoffe, Moussonstrasse 18, CH-8044 Zürich, Switzerland.



Wyszukiwarka

Podobne podstrony:
Rozpiska Mikrobiologia Przemyslowa 09, BIO, PCR - DGGE, In Situ, API, Lab 1
Raki in situ, Dermatologia
Fluorescent proteins as a toolkit for in vivo imaging 2005 Trends in Biotechnology
5 Techniki in situ
Prezentacja-PCR-DGGE, Mikrobiologia, Mikrobiologia, PCR - DGGE, In Situ, API
Prezentacja - PCR-DGGE, Mikrobiologia, Mikrobiologia, PCR - DGGE, In Situ, API
Regulamin zajec Mikrobiologia Przemyslowa 09, BIO, PCR - DGGE, In Situ, API, Lab 1
Stany przedrakowe i raki in situ(2), dermatologia
zad 4?dania in situ prowadzone na obiekcie zabytkowym
13. TERENOWE - in situ, Inżynieria środowiska, inż, Semestr IV, Mechanika gruntów, laboratorium
Osoby prowadzace zajecia Mikrobiologia Przemyslowa 09, BIO, PCR - DGGE, In Situ, API, Lab 1
rak in situ, Download, - ▧ Normalne, higiena
Ca in situ
Wstep, BIO, PCR - DGGE, In Situ, API, Lab 4
Metody in situ i ex situ w bioremediacji
Robert Rix The Barbarian North in Medieval Imagination

więcej podobnych podstron