ArticlePDF Available

A physical vapor deposition method for controlled evaluation of biological response to biomaterial chemistry and topography

Authors:

Abstract and Figures

The purpose of this study was to characterize a technique to effectively mask surface chemistry without modifying surface topography. A thin layer of titanium was deposited by physical vapor deposition (PVD) onto different biomaterial surfaces. Commercially pure titanium disks were equally divided into three groups. Disks were either polished to a mirror finish, grit blasted with alumina particles, or grit blasted and subsequently plasma sprayed with a commercial grade of hydroxyapatite (HA). A subgroup of each of these treatment types was further treated by masking the entire disk surface with a thin layer of commercially pure titanium deposited by PVD. A comparison of surface topography and chemical composition was carried out between disks within each treatment group. Canine marrow cells were seeded on all disk surfaces to determine the stability of the PVD Ti mask under culture conditions. The PVD process did not significantly alter the surface topography of any samples. The thin titanium layer completely masked the underlying chemistry of the plasma sprayed HA surface and the chemistry of the plasma vapor deposited titanium layer did not differ from that of the commercially pure titanium disks. Aliquots obtained from the media during culture did not indicate any significant differences in Ti concentration amongst the Ti and Ti-masked surfaces. The PVD application of a Ti layer on HA coatings formed a stable, durable, and homogenous layer that effectively masked the underlying surface chemistry without altering the surface topography.
Content may be subject to copyright.
A physical vapor deposition method for controlled
evaluation of biological response to biomaterial
chemistry and topography
S.A. Hacking,
1,2,3
M. Zuraw,
4
E.J. Harvey,
1,2,5
M. Tanzer,
1,2,5
J.J. Krygier,
1
J.D. Bobyn
1,2,3,5
1
Jo Miller Orthopaedic Research Laboratory, McGill University, Montreal, Canada
2
Division of Orthopaedics, McGill University, Montreal, Canada
3
Department of Biomedical Engineering, McGill University, Montreal, Canada
4
Fused Metals Inc., Georgetown, Canada
5
Department of Surgery, McGill University, Montreal, Canada
Received 13 October 2005; revised 1 September 2006; accepted 28 September 2006
Published online 31 January 2007 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/jbm.a.31131
Abstract: The purpose of this study was to characterize a
technique to effectively mask surface chemistry without
modifying surface topography. A thin layer of titanium was
deposited by physical vapor deposition (PVD) onto differ-
ent biomaterial surfaces. Commercially pure titanium disks
were equally divided into three groups. Disks were either
polished to a mirror finish, grit blasted with alumina par-
ticles, or grit blasted and subsequently plasma sprayed with
a commercial grade of hydroxyapatite (HA). A subgroup of
each of these treatment types was further treated by mask-
ing the entire disk surface with a thin layer of commercially
pure titanium deposited by PVD. A comparison of surface
topography and chemical composition was carried out
between disks within each treatment group. Canine marrow
cells were seeded on all disk surfaces to determine the sta-
bility of the PVD Ti mask under culture conditions. The
PVD process did not significantly alter the surface topogra-
phy of any samples. The thin titanium layer completely
masked the underlying chemistry of the plasma sprayed
HA surface and the chemistry of the plasma vapor depos-
ited titanium layer did not differ from that of the commer-
cially pure titanium disks. Aliquots obtained from the
media during culture did not indicate any significant differ-
ences in Ti concentration amongst the Ti and Ti-masked
surfaces. The PVD application of a Ti layer on HA coatings
formed a stable, durable, and homogenous layer that effec-
tively masked the underlying surface chemistry without
altering the surface topography. Ó2007 Wiley Periodicals,
Inc. J Biomed Mater Res 82A: 179–187, 2007
Key words: hydroxyapatite; surface chemistry; physical
vapor deposition; biomaterial; surface topography
INTRODUCTION
Biomaterial evaluation is becoming increasingly more
complex as new manufacturing techniques, materials,
coatings, methods of analysis, and devices are devel-
oped. The long-term success of any implantable device
depends in part upon its interaction with the host tissue.
In many cases, a better understanding of the tissue–
implant interaction is fundamental to improving implant
function or performance. As a result, the relationship
between the implant substrate and the peri-implant
environment is a subject of great interest.
In such investigations, it is important to recognize
that subtle changes in biomaterial surface morphol-
ogy can have profound effects on peri-implant tissue
formation
1–11
or cellular
7,12–17
response. These responses
to surface topography, or more precisely, surface
morphology, need to be dissociated from other factors
such as local surface chemistry in order to determine
the exact causes for tissue response. As a result, ex-
perimental controls must not just approximate sur-
face morphology, but match it exactly, since subtle
changes in surface morphology have the potential to
confound experimental findings.
However, it may not be practical or even possible
to produce an exact morphological control from a dif-
ferent biomaterial. This approach can be particularly
problematic in cases where the surface morphology is
complex
5,18–20
or is a consequence of the manufactur-
ing process.
5,18–22
One solution is to produce a control
by coating the surface of interest with a thin film that
is dense, durable, homogenous, and does not alter the
surface morphology. To be truly useful, the film must
also be practical to apply and the ideal result of the
application of the thin film would be a change only in
the surface chemistry of the coated component.
Correspondence to: S.A. Hacking; e-mail: ahacking@yahoo.
com
'2007 Wiley Periodicals, Inc.
There are a variety of thin film deposition tech-
niques collectively referred to as physical vapor depo-
sition (PVD): DC and RF magnetron sputtering,
23–25
electron beam (E-Beam) evaporation,
26–28
cathodic arc
and thermal evaporation. Medical applications of PVD
coatings include wear resistant coatings
29–31
or bio-
compatible coatings
29,32–35
for medical devices. While
other coating methods exist, PVD has the advantage
of being a viable process at lower temperatures in a
relatively inert environment.
36
In the PVD process, an
ionized gas (e.g., Ar
þ
) strikes a target (cathode) and
releases the source material which travels across a
vacuum and condenses on the surface to be coated. E-
Beam coatings are typically produced at about
5008C.
36
Cathodic arc coatings are typically deposited
at temperatures much lower than 4008C, and in some
cases near room temperature.
36,37
The hypothesis of this study was that a thin, com-
mercially pure titanium film applied by PVD could
effectively mask the surface chemistry of a textured
biomaterial without substantially altering its surface
topography.
MATERIALS AND METHODS
PVD process (PVD-mask)
A modified PVD process was developed to coat (mask)
a variety of specimens described later with commercially
pure titanium. All samples were placed directly into the
PVD chamber. The base pressure before processing was
not greater than 6 10
2
Pa. Radiant heaters maintained a
temperature of 1408C, as measured in the center of the ves-
sel. The chamber was backfilled with prepurified argon
(99.99% pure) to a pressure of 2.0 Pa. Titanium was evapo-
rated from a commercially pure source (99.99%) with an
arc current of 125 A. A bias voltage of 250 V pulsing at a
frequency of 20 kHz was maintained on the specimens.
Initial experiments with this technique had shown that
small particles of Ti condensed on the specimen surface
[Fig. 1(A)]. To reduce this contamination, a shield with a
floating bias with respect to ground was placed 10 cm in
front of the titanium source [Fig. 1(B)].
Test specimens
One hundred and twenty cell culture disks, 22 mm in
diameter and 3 mm thick, were divided equally into three
groups. One group was polished (Pol), one group was grit
blasted (GB) with Al
2
O
3
particles, and another group was
grit blasted then plasma sprayed with a layer of hydroxy-
apatite (HA) using industry standard techniques (Implex
Corp, Allendale, NJ) (Fig. 2). The plasma spray process
resulted in a 60-mm thick HA coating that was 98% HA
and 64% crystalline with a density of 99% and a calcium:
phosphate ratio of 1.67. Twenty of the Pol, GB, and HA
disks were left untreated and 20 of each group were PVD-
masked, as described earlier. Representative disks from
each group were randomly selected for surface characteri-
zation.
Characterization of the disk surfaces
Implant topography
Scanning electron micrographs of all disk surfaces were
obtained to provide a qualitative impression of surface
morphology. Surface topography was quantified using a
Wyko NT 2000 (Veeco, Rochester, NY) noncontact optical
profiler. The profiler was calibrated before use and the
operational parameters were: VSI mode, 52Mag, VSI fil-
ter, and tilt correction. Three random regions from three
disks of each group were analyzed yielding nine measure-
ments per surface.
Surface chemistry
Surface chemical analysis was performed using X-ray
photoelectron spectroscopy (XPS). All measurements were
carried out using a dual-anode source in a VG Escalab
MKII instrument (Thermo VG Scientific, Beverly, MA) with
nonmonochromatized Mg Karadiation (hn¼1253.6 eV)
operated at 20 mA and 15 kV. Survey spectra were
Figure 1. (A) First generation PVD Ti surface showing
particle contamination and (B) second generation PVD Ti
surface with a gross reduction in particle contamination.
180 HACKING ET AL.
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
obtained at 908from the sample surface using a pass
energy of 100 eV, 1.0 eV steps, and a 15 mm 6 mm slit-
width, which result in an analyzed surface area of 3 mm
2 mm. When present, specimen-charging effects were com-
pensated by adjusting the binding energy of the survey
spectra to fix the binding energy of the hydrocarbon peak
at 285.0 eV. The concentration of each element was deter-
mined from the XPS signal area and the corresponding
XPS atomic sensitivity factor relative to Fluorine 1s elec-
tron. The sensitivity of the technique was 0.1 at 100%.
With this technique, measurements below 0.2% are consid-
ered contaminant levels. Three cell culture disks from each
group were selected for analysis. Five locations were ran-
domly analyzed on each disk.
XPS was selected for the chemical analysis of the disk
surfaces since it has an effective penetration depth of
25 nm. This was especially necessary for the PVD-mask
specimens to limit analysis to the thin PVD Ti film without
influence from the substrate below.
The Pol and GB disks were scanned with XPS and com-
pared with the PVD-mask, Pol, and GB disks. Scans of the
PVD-mask HA-coated disks were also included to deter-
mine if the underlying HA Ca and P chemistry could be
detected. Three disks were randomly selected from each
group and scanned five times each.
Stability of the PVD coating in an in vitro
cell culture model
An in vitro cell culture model was utilized to determine
the effect of tissue growth and culture media on the control
and PVD-mask surfaces (n¼4). Canine marrow cells
obtained from the iliac crest were harvested in a sterile envi-
ronment, maintained in a-MEM medium supplemented
with 10% FBS, 100 U/mL penicillin, and 100 mg/mL strepto-
mycinandgrowninprimarycultureforaperiodof7days.
Cells were seeded either on disks or on tissue culture plastic
(TCP) at a plating density of 3 10
5
cells/cm
2
(100,000 cells
per disk) in 12-well plates and maintained in 2.0 mL stand-
ard medium (a-MEM containing 10% FBS, 100 U/mL peni-
cillin, and 100 mg/mL streptomycin and 50 mg/mL ascorbate
and 5 mMb-glycerophosphate) in a 5% CO
2
air-balanced in-
cubator at 378Cuptoday12.
Media was changed every 2 days and aspirates were
obtained prior to media change at days 0, 2, 5, 7, 10, and
12 days. The resistance to dissolution of the PVD-mask
film used in this experiment was determined by analysis
of the Ti concentration in aliquots of media from cell cul-
ture. Aliquots from the HA-coated and PVD-mask HA-
coated disks were also analyzed for Ca concentration. Ali-
quots were analyzed by a sequential inductively coupled
plasma spectrometer (Trace Scan, Jarrell-Ash Corp, Frank-
lin, MA). Certified commercial standards (1000 ppm) were
used and subsequently diluted with deionized water.
Standard concentrations bracketed the test samples. Two-
level standardization was used and titanium compounds
were used in the standard preparation. The detection lim-
its for titanium and calcium by this system were 5 ppb
and 5 ppm, respectively.
Assessment of cell proliferation
A parallel set of disks (n¼4 at each time period) was
harvested at days 2, 5, 7, and 12 to demonstrate cell
Figure 2. Cell culture disks used in the study. (A) Polished Ti, (B) polished Ti þPVD mask, (C) grit blasted Ti, (D) grit
blasted Ti þPVD mask, (E) HA coated, and (F) HA coated þPVD mask. Small slot in disk facilitates media change with-
out disturbing culture surface.
PVD METHOD FOR MASKING SURFACE CHEMISTRY 181
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
growth and proliferation. Cell proliferation was deter-
mined by quantification of total DNA according to the
method of Labarca and Paigen.
38
Briefly, culture media
was aspirated from the wells and disks were washed three
times in PBS, following which cells were harvested in a so-
lution of 2MPBS. In cases where a cell layer existed it was
removed in its entirety, otherwise cells were removed by a
combination of cell scraping and vigorous pipetting. Previ-
ous experiments determined that this method provided
more consistent results than using trypsin or collagenase.
Cells were kept on ice and lysed by sonification (2 30 s).
DNA content was determined by Hoescht dye.
39
The sample solutions were diluted as follows: 1 day –
no dilution, 3 days – 5 times dilution, 6 days – 10 times
dilution, 12 days – 20 times dilution. Each sample received
35 mL of Hoechst dye and if the sample was diluted the
balance to 965 mL was made up with 2MPBS.
Where appropriate data were analyzed with the Stu-
dent’s ttest or with ANOVA and various post hoc tests. All
analysis was performed at the 95% confidence level.
RESULTS
Surface topography
Qualitative SEM images of the surfaces are pre-
sented in Figure 3. Three dimensional scans of the
test surfaces are presented in Figure 4. There was no
significant difference in surface topography between
the Pol (R
a
¼0.01 60.02 mm) and Pol PVD-mask
(R
a
¼0.01 60.04 mm) disks, between the GB (R
a
¼
3.26 60.95 mm) and GB PVD-mask (R
a
¼3.32 61.38
mm) disks or between the HA (R
a
¼2.98 60.81 mm)
and HA PVD-mask (R
a
¼3.01 60.37 mm) disks (Table I).
The preservation of surface topography is further
illustrated in Figure 1(B), where one half of a polished
disk was PVD-masked. The continuity of small pol-
ishing artifacts is visible on both the polished and pol-
ished PVD-mask regions.
Surface chemistry
Chemical analyses of the specimen surfaces are pre-
sented in Figure 5. The chemical profiles for the Pol,
Pol PVD-mask, GB, GB PVD-mask, and HA PVD-
mask surfaces were nearly identical. This indicated
that on all test surfaces the PVD deposited Ti film
possessed a surface chemistry that was the same as
the Ti substrate. As anticipated, no Ca or P was
detected on the Pol, Pol PVD-mask, GB, or GB PVD-
mask disks. Ca and P were detected in a ratio of *2:1
on the HA-coated disks. Importantly, there was no Ca
or P detected in any of the 15 scans of the PVD-mask
HA disks, indicating that the applied Ti film had pro-
duced a homogeneous mask that effectively occluded
the chemistry of the underlying HA surface. This fur-
ther indicated that the thickness of the PVD Ti coating
was at least 25 nm, the maximum penetration depth
of XPS.
In vitro stability of the PVD titanium mask
Cell growth and proliferation occurred on all disks
except the HA-coated disks (Fig. 6). There was no sig-
nificant difference in the rate of cell growth between
Figure 3. Scanning electron micrographs of cell culture disks with (A) polished Ti, (B) polished Ti þPVD Ti mask, (C)
grit blasted Ti, (D) grit blasted Ti þPVD Ti mask, (E) HA coated, (F) HA coated þPVD Ti mask surface. SEM, 1000.
182 HACKING ET AL.
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
Pol Ti and Pol Ti þPVD mask and GB Ti and GB
Ti þPVD mask surfaces.
There was a slight increase in Ti levels (40–50 ppb)
in the culture media of each sample (Table II). These
Ti levels were extremely low and near the detection
limits of the instrument involved (5 ppb). A paired
Student’s ttest showed that there was no significant
difference in Ti levels detected between Pol and Pol
PVD-mask, GB and GB PVD-mask, and HA PVD-
mask disks. The rougher disks (GB and GB PVD-
mask, HA PVD-mask disks) had a slightly higher
level of Ti in the media when compared with the
smooth disks (Pol and Pol PVD-mask), however, this
difference was not significant.
There was a significant difference in Ca level from
aspirates of the HA and HA PVD-mask disks (Ta-
ble III). Calcium aspirate concentration for the HA-
coated disks was maximal at day 2 at 2452 642 ppm
and gradually diminished to 714 636 ppm at day 12.
In contrast, the calcium aspirate concentration for the
HA PVD-mask disks was 62.5 62.1 ppm at day 2
and 58.4 61.9 ppm at day 12. This compared well
with the blank value of 69.8 60.71 ppm.
DISCUSSION
Subtle changes in surface morphology have signifi-
cant and predictable effects on cellular behavior and
peri-implant tissue formation. In the context of bio-
medical investigations, it is imperative that surface
morphology be strictly controlled as an experimental
variable to avoid confounding experimental results.
In this study, a 100-nm Ti film applied by PVD effec-
tively masked the surface chemistry of an HA-coated
implant without altering its surface topography.
Other researchers have used similar techniques
such as RF sputtering to successfully produce thin Ti
films for biomedical evaluation.
40–46
Thin TiO, TiN,
and C films have also been successfully applied to
mask the underlying chemistry to prevent corrosion
or ion release from CoCr alloy surfaces.
27,47–49
How-
ever, as far as the authors are aware, this is the first
study to use a thin Ti film to specifically produce a
morphological control with a different surface chem-
istry.
A comparison of the surface topographies indicates
that the application of the PVD coating did not signif-
icantly alter the surface roughness of any of the cul-
ture surfaces. Chemical analysis of the Pol and GB
disks with and without the PVD mask demonstrated
that the application of the PVD Ti mask did not pro-
duce a surface coating that was significantly different
from that of the uncoated Ti specimens. There were
trace levels of Al present in the POL samples, which
Figure 4. Non contact optical profile of cell culture disks with (A) polished Ti, (B) polished Ti þPVD Ti mask, (C) grit
blasted Ti, (D) grit blasted Ti þPVD Ti mask, (E) HA coated, (F) HA coated þPVD Ti mask surfaces. 102vertical reso-
lution not to scale amongst images.
TABLE I
Surface Roughness of Cell Culture Surfaces (n=9)
Surface
Roughness Parameter (mm)
R
a
R
q
TCP <0.01 <0.01
Pol Ti Disk 0.01 60.02 0.02 60.02
Pol Ti Disk þPVD Ti Mask 0.01 60.03 0.02 60.03
GB Ti Disk 3.26 60.95 4.04 61.04
GB Ti Disk þPVD Ti Mask 3.32 61.38 4.01 61.53
HA-coated Disk 2.98 60.81 3.72 61.03
HA-coated Disk þPVD Ti Mask 3.01 60.37 3.87 60.41
PVD METHOD FOR MASKING SURFACE CHEMISTRY 183
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
were most likely a residue from the Al
2
O
3
polishing
process. Analysis of the HA and PVD-mask HA sam-
ples demonstrated that the PVD Ti mask produced a
uniform and dense barrier that effectively masked the
underlying surface chemistry. Ca and P were detected
in a *2:1 ratio in the HA-coated disk, but were not
detected on the PVD-mask HA-coated disks.
For all samples where cell growth occurred, there
was slight elevation of Ti in the media. The Ti levels
presented here are higher than those of the PVD-
mask HA in the previous soak study (40 ppb vs.
5 ppb), but still in the order of parts per billion. It is
interesting to note that the rougher surfaces (GB Ti,
PVD GB Ti, and PVD HA) had a slightly elevated
level of Ti in the media, compared with the Pol surfa-
ces; this perhaps was due to their greater surface
area. At some time periods, aspirations from the HA-
coated disks indicated trace amounts of Ti, near the
detection limits of the instrument. This Ti artifact may
be a result of the media being exposed to Ti from the
non-HA-coated sides and bottom of the disks (only
the top of the disk was coated with HA).
Figure 5. Chemical analysis of the culture disk surfaces used in this study. There is no Ca or P present in any of the
PVD Ti mask HA samples. In addition, the chemical profile of the Ti and PVD mask Ti samples are not significantly dif-
ferent.
Figure 6. Growth of canine marrow culture on Pol Ti, PVD-masked Pol Ti, GB Ti, PVD-masked GB Ti, HA, PVD-masked
HA, and TCP surfaces. Cells proliferated on all surfaces except the HA surface.
184 HACKING ET AL.
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
A previous soak study also confirmed the stability
of the PVD coating for long periods of time in a modi-
fied saline environment.
50
The aim of the present
study was to test the stability of the PVD coating
under tissue culture conditions. With the exception
the HA-coated disks, all samples exhibited cell
growth and proliferation. In the current study, media
aspirated from the HA-coated disks presented signifi-
cantly increased Ca levels. To determine if cellular ac-
tivity was affecting the Ca levels, a subsequent test
was conducted by immersing the same HA-coated
disks in the same media and incubating for 12 days
without cells. In this soak test, elevated levels of Ca
were observed that were not significantly different to
those reported in this study. This finding suggests
that there may be an interaction between the HA and
culture media. While HA is generally not soluble, the
HA coating used in this study was amorphous (36%
noncrystalline, amorphous HA).
The elevated levels of Ca are believed to arise from
a dissolution of the HA coating. Since the aim of this
study was to determine if the PVD Ti mask effectively
concealed the HA chemistry under culture conditions,
the dramatically elevated Ca levels from aspirates of
the HA-coated surfaces are a fortunate finding. The
levels of Ca in the PVD Ti-masked samples were
lower than the measured media Ca concentration at
day 0 (aspirates were obtained after 2 days of cul-
ture). If the HA coating from the PVD Ti mask HA
disks had been exposed to the media, a much higher
concentration of Ca would have been expected, indi-
cating once again that the PVD-mask effectively con-
cealed the underlying HA surface chemistry.
It is important to note that Ti was selected as the
coating material for this study because of its resist-
ance to degradation in a physiologic environ-
ment.
40,51,52
Other investigators have coated implant
surfaces with thin films of CaP by sputtering techni-
ques.
53–57
However, in this context, the degradation
of these thin CaP coatings with subsequent change in
morphology upon exposure to physiologic fluids or
media mitigates their utility as an effective coating for
long term morphological controls. A plasma spray
HA surface was selected as a substrate material since
it was chemically dissimilar to the commercially pure
Ti mask, but like Ti it is commonly associated with
orthopaedic implants. The plasma spray HA surface
also presented a complex geometry that tested the ef-
ficacy of the PVD Ti to mask irregular biomaterial
substrates.
For nonporous metallic materials such as Ti in non-
abrasive conditions like these, only the first 10 nm of
the substrate surface actually participates in any
chemical interaction with the surrounding environ-
ment.
52
Since the PVD Ti coating used in this study
was 100 nm thick, it is reasonable to assume that the
tissue reaction to the PVD mask Ti-coated implant
was no different from that of the same implant made
entirely of Ti.
In conclusion, the PVD Ti coating effectively
masked the underlying surface chemistry of the HA-
coated disk without altering its surface topography.
In addition, the PVD process produced a thin, dense,
and homogenous Ti film possessing the same surface
chemistry as commercially pure Ti. The PVD of thin
Ti films has utility for the investigation of the effects
TABLE II
Dissolved Titanium in Media Aspirates
Disk Surface
Dissolved Titanium (ppb)
0
a
2 5 7 10 12
Pol Ti 0 38.2 65.6 41.0 65.8 37.3 64.6 39.0 62.0 40.1 63.8
Pol Ti þPVD 0 39.8 67.5 43.8 67.3 41.8 62.8 40.3 63.5 40.3 62.8
GB Ti 0 44.9 63.7 46.9 63.8 47.7 61.9 49.3 62.2 48.1 63.5
GB Ti þPVD 0 46.3 64.4 51.2 65.6 50.5 62.6 48.9 64.6 50.6 62.5
HA 0 2.5 62.9 1.3 62.5 0.0 60 1.8 63.5 1.5 63
HA þPVD 0 45.5 64.4 50.8 61.9 47.1 63.6 46.8 64.9 49.9 64.8
TCP 000000
a
0, 2, 5, 7, 10, and 12 are time periods (days).
TABLE III
Dissolved Calcium in Media Aspirates
Disk Surface
Dissolved Calcium (ppm)
0
a
2571012
HA þPVD 69.8 60.71 62.5 62.1 59.8 61.8 58.1 62.0 58.5 61.8 58.4 61.9
HA 70.2 60.66 2452 642 1913 645 1030 638 844 641 714 636
a
0, 2, 5, 7, 10, and 12 are time periods (days).
PVD METHOD FOR MASKING SURFACE CHEMISTRY 185
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
of surface topography and surface chemistry on the cel-
lular and/or tissue response to various biomaterials.
References
1. Chehroudi B, Gould TR, Brunette DM. A light and electron
microscopic study of the effects of surface topography on the
behavior of cells attached to titanium-coated percutaneous
implants. J Biomed Mater Res 1991;25:387–405.
2. Clark P, Connolly P, Curtis AS, Dow JA, Wilkinson CD. To-
pographical control of cell behaviour. I. Simple step cues. De-
velopment 1987;99:439–448.
3. Clark P, Connolly P, Curtis AS, Dow JA, Wilkinson CD. To-
pographical control of cell behaviour. II. Multiple grooved
substrata. Development 1990;108:635–644.
4. Curtis A, Wilkinson C. Topographical control of cells. Bioma-
terials 1997;18:1573–1583.
5. Curtis A, Wilkinson C. New depths in cell behaviour: Reac-
tions of cells to nanotopography. Biochem Soc Symp 1999;
65:15–26.
6. Curtis AS, Wilkinson CD. Reactions of cells to topography.
J Biomater Sci Polym Ed 1998;9:1313–1329.
7. Keller JC. Tissue compatibility to different surfaces of dental
implants: In vitro studies. Implant Dent 1998;7:331–337.
8. Kieswetter K, Schwartz Z, Dean DD, Boyan BD. The role of
implant surface characteristics in the healing of bone. Crit
Rev Oral Biol Med 1996;7:329–345.
9. Lincks J, Boyan BD, Blanchard CR, Lohmann CH, Liu Y,
Cochran DL, Dean DD, Schwartz Z. Response of MG63 osteo-
blast-like cells to titanium and titanium alloy is dependent on
surface roughness and composition. Biomaterials 1998;19:
2219–2232.
10. Meyle J, Wolburg H, von Recum AF. Surface micromorphol-
ogy and cellular interactions. J Biomater Appl 1993;7:362–374.
11. Park JY, Gemmell CH, Davies JE. Platelet interactions with ti-
tanium: Modulation of platelet activity by surface topogra-
phy. Biomaterials 2001;22:2671–2682.
12. Abrahamsson I, Zitzmann NU, Berglundh T, Linder E,
Wennerberg A, Lindhe J. The mucosal attachment to titanium
implants with different surface characteristics: An experimen-
tal study in dogs. J Clin Periodontol 2002;29:448–455.
13. Bigerelle M, Anselme K, Dufresne E, Hardouin P, Iost A. An
unscaled parameter to measure the order of surfaces: A new
surface elaboration to increase cells adhesion. Biomol Eng
2002;19:79–83.
14. DegasneI,BasleMF,DemaisV,HureG,LesourdM,GrolleauB,
Mercier L, Chappard D. Effects of roughness, fibronectin and vi-
tronectin on attachment, spreading, and proliferation of human
osteoblast-like cells (Saos-2) on titanium surfaces. Calcif Tissue Int
1999;64:499–507.
15. den Braber ET, de Ruijter JE, Ginsel LA, von Recum AF,
Jansen JA. Orientation of ECM protein deposition, fibroblast
cytoskeleton, and attachment complex components on silicone
microgrooved surfaces. J Biomed Mater Res 1998;40:291–300.
16. Dunlap D, Cattelino A, de Curtis I, Valtorta F. Cytoplasmic
topography of focal contacts. FEBS Lett 1996;382:65–72.
17. Locci P, Becchetti E, Pugliese M, Rossi L, Belcastro S, Calvitti
M, Pietrarelli G, Staffolani N. Phenotype expression of
human bone cells cultured on implant substrates. Cell Bio-
chem Funct 1997;15:163–170.
18. Bobyn JD, Stackpool GJ, Hacking SA, Tanzer M, Krygier JJ.
Characteristics of bone ingrowth and interface mechanics of a
new porous tantalum biomaterial. J Bone Joint Surg Br
1999;81:907–914.
19. Clark P, Connolly P, Curtis AS, Dow JA, Wilkinson CD. Cell
guidance by ultrafine topography in vitro. J Cell Sci 1991;99(Part
1):73–77.
20. Duncan AC, Weisbuch F, Rouais F, Lazare S, Baquey C. Laser
microfabricated model surfaces for controlled cell growth.
Biosens Bioelectron 2002;17:413–426.
21. Hacking SA, Bobyn JD, Tanzer M, Krygier JJ. The osseous
response to corundum blasted implant surfaces in a canine
hip model. Clin Orthop Relat Res 1999;364:240–253.
22. Wong M, Eulenberger J, Schenk R, Hunziker E. Effect of sur-
face topology on the osseointegration of implant materials in
trabecular bone. J Biomed Mater Res 1995;29:1567–1575.
23. Hulshoff JE, Hayakawa T, van DK, Leijdekkers-Govers AF,
van der Waerden JP, Jansen JA. Mechanical and histologic
evaluation of Ca-P plasma-spray and magnetron sputter-
coated implants in trabecular bone of the goat. J Biomed
Mater Res 1997;36:75–83.
24. Vercaigne S, Wolke JG, Naert I, Jansen JA. A histological
evaluation of TiO
2
-gritblasted and Ca-P magnetron sputter
coated implants placed into the trabecular bone of the goat,
Part 2. Clin Oral Implants Res 2000;11:314–324.
25. Wolke JG, van der Waerden JP, de GK, Jansen JA. Stability of
radiofrequency magnetron sputtered calcium phosphate coat-
ings under cyclically loaded conditions. Biomaterials 1997;18:
483–488.
26. Nan H, Ping Y, Xuan C, Yongxang L, Xiaolan Z, Guangjun C,
Zihong Z, Feng Z, Yuanru C, Xianghuai L, Tingfei X. Blood com-
patibility of amorphous titanium oxide films synthesized by ion
beam enhanced deposition. Biomaterials 1998;19:771–776.
27. Pan J, Leygraf C, Thierry D, Ektessabi AM. Corrosion resist-
ance for biomaterial applications of TiO
2
films deposited on
titanium and stainless steel by ion-beam-assisted sputtering.
J Biomed Mater Res 1997;35:309–318.
28. Zhang F, Zheng Z, Chen Y, Liu X, Chen A, Jiang Z. In vivo
investigation of blood compatibility of titanium oxide films.
J Biomed Mater Res 1998;42:128–133.
29. Arweiler-Harbeck D, Sanders A, Held M, Jerman M, Ehrich
H, Jahnke K. Does metal coating improve the durability of
silicone voice prostheses? Acta Otolaryngol 2001;121:643–646.
30. Ruddell DE, Thompson JY, Stoner BR. Mechanical properties
of a dental ceramic coated by RF magnetron sputtering.
J Biomed Mater Res 2000;51:316–320.
31. Ward LP, Subramanian C, Strafford KN, Wilks TP. Sliding
wear studies of selected nitride coatings and their potential
for long-term use in orthopaedic applications. Proc Inst Mech
Eng [H] 1998;212:303–315.
32. Auditore A, Satriano C, Coscia U, Ambrosone G, Parisi V,
Marletta G. Human serum albumin adsorption onto a-SiC:H
and a-C:H thin films deposited by plasma enhanced chemical
vapor deposition. Biomol Eng 2002;19:85.
33. Bobyn JD, Toh KK, Hacking SA, Tanzer M, Krygier JJ. Tissue
response to porous tantalum acetabular cups: A canine model.
J Arthroplasty 1999;14:347–354.
34. Hacking SA, Bobyn JD, Toh K, Tanzer M, Krygier JJ. Fibrous
tissue ingrowth and attachment to porous tantalum. J Biomed
Mater Res 2000;52:631–638.
35. Mohanty M, Anilkumar T, Mohanan P, Muraleedharan C,
Bhuvaneshwar G, Derangere F, Sampeur Y, Suryanarayanan
R. Long term tissue response to titanium coated with dia-
mond like carbon. Biomol Eng 2002;19:125–128.
36. Staia MH, Lewis B, Cawley J, Hudson T. Chemical vapour
deposition of TiN on stainless steel. Surf Coat Technol 1995;
76/77:231–236.
37. Treglio JR, Tian AF, Perry AJ. Low-temperature deposition of
titanium nitride. Surf Coat Technol 1995;76/77:815–820.
38. Labarca C, Paigen K. A simple, rapid, and sensitive DNA
assay procedure. Anal Biochem 1980;102:344–352.
39. Araki T, Yamamoto A, Yamada M. Accurate determination of
DNA content in single cell nuclei stained with Hoechst 33258
fluorochrome at high salt concentration. Histochemistry 1987;
87:331–338.
186 HACKING ET AL.
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
40. Albrektsson T, Hansson HA. An ultrastructural characteri-
zation of the interface between bone and sputtered tita-
nium or stainless steel surfaces. Biomaterials 1986;7:201–
205.
41. Boyan BD, Schwartz Z, Hambleton JC. Response of bone and
cartilage cells to biomaterials in vivo and in vitro. J Oral
Implantol 1993;19:116–122.
42. Hambleton J, Schwartz Z, Khare A, Windeler SW, Luna M,
Brooks BP, Dean DD, Boyan BD. Culture surfaces coated
with various implant materials affect chondrocyte growth
and metabolism. J Orthop Res 1994;12:542–552.
43. Kemmenoe BH, Bullock GR. Structure analysis of sputter-
coated and ion-beam sputter-coated films: A comparative
study. J Microsc 1983;132(Part 2):153–163.
44. Koontz CS, Ramp WK, Peindl RD, Kaysinger KK, Harrow
ME. Comparison of growth and metabolism of avian osteo-
blasts on polished disks versus thin films of titanium alloy.
J Biomed Mater Res 1998;42:238–244.
45. Mezger PR, Creugers NH. Titanium nitride coatings in clini-
cal dentistry. J Dent 1992;20:342–344.
46. Schwartz Z, Kieswetter K, Dean DD, Boyan BD. Underlying
mechanisms at the bone-surface interface during regenera-
tion. J Periodontal Res 1997;32:166–171.
47. Hendry JA, Pilliar RM. The fretting corrosion resistance of
PVD surface-modified orthopedic implant alloys. J Biomed
Mater Res 2001;58:156–166.
48. Lappalainen R, Anttila A, Heinonen H. Diamond coated
total hip replacements. Clin Orthop Relat Res 1998;352:118–
127.
49. Wisbey A, Gregson PJ, Tuke M. Application of PVD TiN
coating to Co-Cr-Mo based surgical implants. Biomaterials
1987;8:477–480.
50. Hacking SA, Tanzer M, Harvey EJ, Krygier JJ, Bobyn JD. Rel-
ative contributions of chemistry and topography to the os-
seointegration of hydroxyapatite coatings. Clin Orthop Relat
Res 2002;405:24–38.
51. Branemark PI, Adell R, Breine U, Hansson BO, Lindstrom J,
Ohlsson A. Intra-osseous anchorage of dental prostheses. I. Ex-
perimental studies. Scand J Plast Reconstr Surg 1969;3:81–100.
52. Kasemo B. Biocompatibility of titanium implants: Surface sci-
ence aspects. J Prosthet Dent 1983;49:832–837.
53. Ding SJ, Ju CP, Lin JH. Characterization of hydroxyapatite
and titanium coatings sputtered on Ti-6Al-4V substrate.
J Biomed Mater Res 1999;44:266–279.
54. Massaro C, Baker MA, Cosentino F, Ramires PA, Klose S,
Milella E. Surface and biological evaluation of hydroxyapa-
tite-based coatings on titanium deposited by different techni-
ques. J Biomed Mater Res 2001;58:651–657.
55. Ozeki K, Yuhta T, Aoki H, Nishimura I, Fukui Y. Push-out
strength of hydroxyapatite coated by sputtering technique in
bone. Biomed Mater Eng 2001;11:63–68.
56. van Dijk K, Maree CH, Verhoeven J, Habraken FH, Jansen
JA. A complete characterization of Ca
5
(PO
4
)3OH sputter-de-
posited films by ion beam analysis: RBS and ERD. J Biomed
Mater Res 1998;42:266–271.
57. Wang CX, Chen ZQ, Wang M, Liu ZY, Wang PL. Ion-beam-sput-
tering/mixing deposition of calcium phosphate coatings. I. Effects
of ion-mixing beams. J Biomed Mater Res 2001;55:587–595.
PVD METHOD FOR MASKING SURFACE CHEMISTRY 187
Journal of Biomedical Materials Research Part A DOI 10.1002/jbm.a
... To improve the surface properties of titanium and its alloys, various mechanical, chemical, and physical techniques such as shot peening [23], ultrasonic peening [24], laser peening [25], anodization [26], grinding [27], physical vapour deposition [28], and die-sinking electrical discharge [29] are used. Surface treatment techniques for titanium and its alloys offer both benefits and drawbacks. ...
... To improve the surface properties of titanium and its alloys, various mechanical, chemical, and physical techniques such as shot peening [23], ultrasonic peening [24], laser peening [25], anodization [26], grinding [27], physical vapour deposition [28], and die-sinking electrical discharge [29] are used. Surface treatment techniques for titanium and its alloys offer both benefits and drawbacks. ...
Article
Full-text available
The process of machining micro surface patterns on a workpiece to improve various performance aspects of engineering materials, including wear resistance, corrosion resistance, and biocompatibility, has been a hot topic of research in recent years. Due to the restricted machinability of titanium and its alloys, it is very challenging to process micro surface patterns with exact surface geometries using traditional machining methods. Consequently, non-traditional processing techniques, such as laser, electro-erosion, and chemical etching, may overcome these obstacles. In the present study, electrical discharge machining (EDM) is used to form micro surface patterns on Cp-Ti alloy samples. First, graphite electrodes with several channels were manufactured, and then square-shaped surface patterns were processed onto Cp-Ti samples using EDM. To evaluate the machining performance of the process and surface features of the obtained micro surface patterns, the surface morphology and topography of the processed samples were investigated by scanning electron microscopy (SEM) and three-dimensional (3D) optical profilometry, respectively. The average widths of the square-shaped surface patterns along the X and Y axes were 663.7±8 µm and 609.5±4 µm, respectively. For micro surface designs with square geometry, dimensional consistency was obtained with exceedingly small amounts of variation. However, a limited number of microcracks were observed due to rapid cooling during the processing of the surface patterns. The 3D surface topographies revealed that square-shaped micro surface patterns were successfully processed on the samples, indicating that micro surface patterns can be processed on Cp-Ti samples by using the proposed methodology, which has the potential for obtaining tailor-designed surface features, particularly for biomedical and tribological applications.
... These physical vapour deposition (PVD) methods can be also applied for obtaining biocompatible coatings, since PVD is a viable process at lower temperatures in a relatively inert environment. In this process, an ionized gas (e.g., Ar + ) strikes a target (cathode) and releases the target material which travels across a vacuum to the substrate making coating on its surface [24,25]. On their way to the substrate, the sputtered atoms can collide with gas atoms which leads to their energy loss. ...
Article
Thin films on the base of titanium oxynitrides and copper doped titanium nitride were obtained by a combination of different PVD techniques on the surface of glass substrate and their physicochemical properties were analyzed in details. Phase composition of the samples was analyzed by XRD and FTIR methods, while microstructure of the samples was analyzed by SEM. XPS was used for depth profiling of the samples, which enabled determination of the oxidative state of titanium and corresponding phases through various film layers from the surface to the substrate. The depth of the various layers and their extinction coefficients and refractory indexes were estimated by spectroscopic ellipsometry.
... ALD coats materials with atomic-scale precision. This surface modification is about self-limiting chemical reactions on surfaces, therefore, yielding atomic-level control over the film thickness and composition without the need for line-of-site access to the precursor source, for example, Al Oxide (Al 2 O 3 ), TiO 2 , Zn, tin (Sn), Zn oxide (ZnO) or Sn oxide (SnO 2 ) application [1], [5]. The study aimed to assess the physicochemical and mechanical properties of the modified Ti13Nb13Zr alloy, considered to be a material with a high level of biocompatibility and Young's modulus of closer value to bone tissue, which is particularly important in the context of treating the skeletal system and the process of osseointegration. ...
Article
The constantly growing need for the use of implants in osteotomy is mainly due to the aging population and the need for long-term use of this type of biomaterials. Improving implant materials requires the selection of appropriate functional properties. Currently used titanium (Ti) alloys, such as Ti6Al4V and Ti6Al7Nb, are being replaced by materials with better biocompatibility, such as vanadium (V) or niobium (Nb), allowing for creation of the so-called new generation alloys. These new alloys, with the incorporation of zirconium (Zr), iron, and tantalum, possess Young’s modulus close to that of a bone, which further improves the improves the biomaterial’s. biocompatibility. This article describes the atomic layer deposition (ALD) method and its possible applications in the new generation of titanium alloys for biomedical applications. Also, the exemplary results of tin oxide (SnO2) thin coatings deposited by ALD and physical vapor deposition (PVD) methods are presented. This study aimed to evaluate the physicochemical properties of a Ti13Nb13Zr alloy used for elements in the skeletal system. As the temperature and the number of cycles vary, the results demonstrate that the surface area of the samples changes. The uncoated Ti13Nb13Zr alloy exhibits hydrophilic properties. However, all coated specimens improve in this respect and provide improved clinical results. after the applied modification, the samples have a smaller contact angle, but still remain in the range of 0–90°, which makes it possible to conclude that their nature remains hydrophilic. Coating the specimens decreased the mineralization risk of postoperative complications. As a result, the biomaterials demonstrated improved effectiveness, decreased complication indicators, and improved patient well-being
... Similarly, Hacking et al. [35] described a method for successfully concealing surface chemistry without changing the contour of the substrate. In this experiment, PVD was used to create a thin coating of Ti upon several biomaterial substrates. ...
Article
Full-text available
Metallic materials are among the most crucial engineering materials widely utilized as biomaterials owing to their significant thermal conductivity, mechanical characteristics, and bio-compatibility. Although these metallic biomedical implants, such as stainless steel, gold, silver, dental amalgams, Co-Cr, and Ti alloys, are generally used for bone tissue regeneration and repairing bodily tissue, the need for innovative technologies is required owing to the sensitivity of medical applications and to avoid any potential harmful reactions, thereby improving the implant to bone integration and prohibiting infection lea by corrosion and excessive stress. Taking this into consideration , several research and developments in biomaterial surface modification are geared toward resolving these issues in bone-related medical therapies/implants offering a substantial influence on cell adherence, increasing the longevity of the implant and rejuvenation along with the expansion in cell and molecular biology expertise. The primary objective of this review is to reaffirm the significance of surface modification of biomedical implants by enlightening numerous significant physical surface modifications, including ultrasonic nanocrystal surface modification, thermal spraying, ion implantation, glow discharge plasma, electrophoretic deposition, and physical vapor deposition. Furthermore, we also focused on the characteristics of some commonly used biomedical alloys, such as stainless steel, Co-Cr, and Ti alloys.
... These physical vapour deposition (PVD) methods can be also applied for obtaining biocompatible coatings, since PVD is a viable process at lower temperatures in a relatively inert environment. In this process, an ionized gas (e.g., Ar + ) strikes a target (cathode) and releases the target material which travels across a vacuum to the substrate making coating on its surface [24,25]. On their way to the substrate, the sputtered atoms can collide with gas atoms which leads to their energy loss. ...
Article
Thin films on the base of titanium oxynitrides and copper doped titanium nitride were obtained by a combination of different PVD techniques on the surface of glass substrate and their physicochemical properties were analyzed in details. Phase composition of the samples was analyzed by XRD and FTIR methods, while microstructure of the samples was analyzed by SEM. XPS was used for depth profiling of the samples, which enabled determination of the oxidative state of titanium and corresponding phases through various film layers from the surface to the substrate. The depth of the various layers and their extinction coefficients and refractory indexes were estimated by spectroscopic ellipsometry.
... A variety of physical methods have been used to generate bioactive nano-topographies on metal surfaces [6]. For example, physical vapor deposition (PVD) has been applied to create different nano-topographical coatings on Ti surfaces, which have been used to evaluate the effects of surface chemistry and topography on the cellular and/or tissue response [7,8]. PVD processes include evaporation, sputtering and ion plating. ...
Article
Full-text available
The influence of commercially pure titanium (Cp Ti) coatings on the biocompatibility of dental implants using a combination of chemical etching and physical vapor deposition methods were studied. The implants prepared and divided into three groups: without modification (control), surface modified with Ti by thermal evaporation (thermally treated), and surfaces chemically etched with H 2 SO 4 and H 2 O 2 before coating with Ti by thermal evaporation deposition (combination treated). For this in vitro experiments, the surfaces characterized by scanning electron microscopy, X-ray diffraction, atomic force microscopy, thickness measurements and microscopy examinations. While, for these in vivo investigations, the implants inserted into the tibia of New Zealand rabbits. A biomechanical tests and histological analysis were performed to understand the bone-implant interface and torque resistance of the implants. The results show that the average removal torque gradually increases and is highest in the combination treated group. In addition, the histological analysis showed improved quality of bone in response to surface nano-modification; the combination treated implants revealed a well-developed mature bone, characterized by bony threads and haversian canal.
Conference Paper
Full-text available
Acanthamoeba are amoebas that are widely found in nature. Some species of Acanthamoeba are qualified as pathogens and, as a result, can infect humans in various ways. Acanthamoeba Keratitis, which leads to blindness, can result in Granulomatous Amoebic Encephalitis with a fatal course in the brain and secondarily different infections. During these infections, it has to fight macrophages in phagocytosis or with Reactive Oxygen Species (ROS) produced by inflammation. ROS is involved in cell signaling and homeostasis. The iron superoxide dismutase present in Acanthamoeba both protects itself from oxidative stress and the immune effector cells of the host may play a role in the survival of Acanthamoeba by detoxifying the oxidative killing of the parasite. In recent studies, it has been reported that SOD activity is high in the pathogen Acanthamoeba. It is known that the drugs used to treat Acanthamoeba are still not very effective. In this study, antimicrobial effective compounds were investigated in line with the studies included in the PubMed database. The interactions between these compounds and FeSOD (Iron superoxide dismutase), which is thought to play a role in the pathogenicity of Acanthamoeba, were studied in silico. For this purpose, 25 compounds were downloaded from the PubChem database and molecular docking was performed with FeSOD (PDB ID: 6J55), Autodock Vina embedded Chimera 1.15 software. Accordingly, the 3 compounds with the lowest energy and the highest scores were taken as a basis. The compounds with the best binding are, Ellagic acid, Carnosol and Episorosmanol respectively. These compounds were also compared with the positive control of Chlorhexidine. According to the results of In silico, these compounds are promising in terms of the potential to be converted into drugs. However, more detailed studies of these compounds, in vitro and in vivo, are needed.
Article
Full-text available
We have studied the characteristics of bone ingrowth of a new porous tantalum biomaterial in a simple transcortical canine model using cylindrical implants 5 × 10 mm in size. The material was 75% to 80% porous by volume and had a repeating arrangement of slender interconnecting struts which formed a regular array of dodecahedron-shaped pores. We performed histological studies on two types of material, one with a smaller pore size averaging 430 μm at 4, 16 and 52 weeks and the other with a larger pore size averaging 650 μm at 2, 3, 4, 16 and 52 weeks. Mechanical push-out tests at 4 and 16 weeks were used to assess the shear strength of the bone-implant interface on implants of the smaller pore size. The extent of filling of the pores of the tantalum material with new bone increased from 13% at two weeks to between 42% and 53% at four weeks. By 16 and 52 weeks the average extent of bone ingrowth ranged from 63% to 80%. The tissue response to the small and large pore sizes was similar, with regions of contact between bone and implant increasing with time and with evidence of Haversian remodelling within the pores at later periods. Mechanical tests at four weeks indicated a minimum shear fixation strength of 18.5 MPa, substantially higher than has been obtained with other porous materials with less volumetric porosity. This porous tantalum biomaterial has desirable characteristics for bone ingrowth; further studies are warranted to ascertain its potential for clinical reconstructive orthopaedics.
Article
Full-text available
This study evaluated the osseous tissue response to a noncemented metal-backed acetabular component made of a new porous tantalum biomaterial. Eleven dogs with bilateral total hip arthroplasties (22 acetabular implants) were studied for a period of 6 months. Thin section histology, high-resolution radiography, and backscattered scanning electron microscopy revealed that all 22 implants had stable bone-implant interfaces. Regions of bone ingrowth were present in all histologic sections. The depth of bone ingrowth varied from 0.2 mm to the maximal limit of 2 mm. Analyzing contiguous regions of interest across the full bone-implant interface, the mean bone ingrowth for all sections was 16.8% ± 5.7%. In the peripheral regions of the cup where bone-implant contact was most consistent, bone ingrowth averaged 25.1% ± 10.1%. The data indicate that the porous tantalum material is effective for biologic fixation in the dog and may provide a suitable alternative to other porous materials used in acetabular cup design.
Article
Full-text available
This study determined the soft tissue attachment strength and extent of ingrowth to a porous tantalum biomaterial. Eight dorsal subcutaneous implants (in two dogs) were evaluated at 4, 8, and 16 weeks. Upon retrieval, all implants were surrounded completely by adherent soft tissue. Implants were harvested with a tissue flap on the cutaneous aspect and peel tested in a servo-hydraulic tensile test machine at a rate of 5 mm/min. Following testing, implants were dehydrated in a solution of basic fuschin, defatted, embedded in methylmethacrylate, and processed for thin-section histology. At 4, 8, and 16 weeks, the attachment strength to porous tantalum was 61, 71, and 89 g/mm respectively. Histologic analysis showed complete tissue ingrowth throughout the porous tantalum implant. Blood vessels were visible at the interface of and within the porous tantalum material. Tissue maturity and vascularity increased with time. The tissue attachment strength to porous tantalum was three- to six-fold greater than was reported in a similar study with porous beads. This study demonstrated that porous tantalum permits rapid ingrowth of vascularized soft tissue, and attains soft tissue attachment strengths greater than with porous beads. © 2000 John Wiley & Sons, Inc. J Biomed Mater Res, 52, 631–638, 2000.
Article
Voice prostheses, which are used for voice rehabilitation in cancer patients after laryngectomy, usually become colonized with a mixed biofilm of bacteria and Candida after 2-4 months and lose their efficiency. It is essential to ensure the stability and biocompatibility of these implants. With the aid of surface frame analysis we have shown that local antifungal treatment is inadequate for eliminating the deep infiltration and encapsulation of Candida colonies in silicone. A surface that prevents the adhesion of microorganisms is required. Because of its special properties there are few methods available for coating silicone. We employed, for the first time, a new method of surface modification using anodic vacuum arc coating. Using this method it was possible to obtain a solid film of gold or titanium metal with a layer thickness < 100 nm. Resistance against Candida colonization and destruction of coated prostheses were tested both in vitro and in vivo. A titanium coating seemed to provide the optima...
Article
A new technique for depositing hard, dense, well-adhering TiN with excellent adhesion is described. Taking advantage of the high degree of ionization of the material emitted from a cathodic arc source, a short, very high voltage pulse (the order of 5–20 kV for 1–3 μs at a frequency of some 1–2 kHz) is applied to the substrate in addition to the usual d.c. bias during reactive deposition. The high bias accelerates the ions located within the sheath during the short pulse and their momentum modifies the properties of the coatings as indicated above. The technology, termed Hyper-Ion, is readily retrofitted to existing physical vapor deposition systems equipped with cathodic arc sources. The present work presents results on titanium nitride deposited onto temperature-sensitive materials including low alloy steel, and aluminum 6061-T6 at a substrate temperature of 150 °C. The coating has a type-T microstructure and the substrates do not lose mechanical strength.
Article
A study has been conducted regarding the deposition of titanium nitride on stainless steel for potential uses in surgical applications. Titanium nitride was deposited from a gaseous mixture of TiCl4, H2 and N2 using a CVD process at atmospheric pressure and a temperature of 1173 K employing a horizontal quartz reactor and linear flow velocities in the range 20 to 60 cm min−1. Adhesion of the coatings was measured by using the scratch adhesion test and their thickness by using the ball cratering method. The results have shown that at low linear flow velocities the deposits obtained are thin, inhomogeneous and have a poor adhesion to the substrate. In contrast, at high velocities (50 cm min−1) the coatings had better properties and characteristics.
Article
A series of thin (<10 μm), single-layered HA/Ti coatings were deposited on Ti-6Al-4V substrate using a radio frequency magnetron-assisted sputtering system. The adhesion strength, microstructure, and chemistry of the coatings were characterized. Experimental results showed that higher Ti contents in targets or coatings resulted in higher deposition rates. When Ti was added the highly crystalline structure of monolithic HA coating was largely disrupted and the coating became amorphous-like. The highly crystalline structure of the monolithic Ti coating was also disrupted by introducing small amounts of Ca, P, and O into the coating. The HA/Ti coatings had quite uniform thicknesses and appeared smooth, dense, and well bonded to the substrate. A scanning electron microscope with an energy dispersive spectroscopy system showed that monolithic HA, 95HA/5Ti, 25HA/75Ti, and 50HA/50Ti coatings had the lowest Ca/P ratios while the 75HA/25Ti coating had the highest. The adhesion strengths of all coatings were between 60 and 80 MPa. © 1999 John Wiley & Sons, Inc. J Biomed Mater Res, 44, 266–279, 1999.