Вы находитесь на странице: 1из 7

Biomaterials 26 (2005) 73507356

Corrosion behavior of titanium in the presence of


calcium phosphate and serum proteins
Xiaoliang Cheng, Sharon G. Roscoe

Department of Chemistry, Acadia University, Wolfville, Nova Scotia, Canada, B4P 2R6
Available online 14 July 2005
Abstract
The effect of calcium phosphate surface deposit and the surface adsorption of the serum proteins, bovine serum albumin (BSA)
and brinogen, on the corrosion resistance and electrochemical behavior of (cp)titanium in phosphate buffer saline solution (pH 7.4)
was investigated at physiological temperature, 37 1C, using electrochemical impedance spectroscopy and dc electrochemical
polarization techniques. The formation of calcium phosphate deposit on the Ti surface decreased both the corrosion rate at the open
circuit potential (OCP) and the anodic reaction current in the high anodic potential range (42.6 V). Addition of BSA signicantly
moved the OCP towards a more negative (cathodic) potential and inhibited the cathodic corrosion reaction, but did not signicantly
change the corrosion resistance at the OCP. Addition of brinogen showed a similar, but less pronounced effect than BSA. The
possible mechanisms leading to these observed effects are discussed.
r 2005 Elsevier Ltd. All rights reserved.
Keywords: Titanium; Calcium phosphate; Bovine serum albumin; Fibrinogen; Corrosion; Biocompatibility; Electrochemistry
1. Introduction
Titanium and its alloys have been used in implants in
the human body for many years. Although they have
better corrosion resistance and biocompatibility com-
pared with other implant materials (e.g., stainless steels),
their corrosion due to dissolution of Ti and alloying
elements (e.g., V, Zr, Al, etc.) is still a concern, because
the metal ions released into the surrounding tissue may
induce the release of potentially osteolytic cytokines
involved in implant loosening [1], cause discolouration
of the tissue, and even more severe problems such as
inammatory reaction of the tissue [2]. As a result, many
publications have dealt with the corrosion resistance of
titanium implants under various conditions [311].
The human body contains uid saturated in both
calcium and phosphate ions. Similar concentrations of
these ions in a simulated body uid produced a calcium
phosphate layer on the surface of titanium [12,13].
Many investigations have been made on the effect of
calcium phosphate or hydroxyapatite coatings on the
corrosion resistance of titanium and its alloys [1417].
Metikos-Hukovic et al. [14] found that well-crystallized
hydroxyapatite coatings derived from solgel exhibited a
benecial corrosion protection effect on the titanium
substrate during prolonged exposure to HBSS. Lavos-
Valereto et al. [15] investigated the electrochemical
behavior of Ti6Al7Nb alloy with and without plasma-
sprayed hydroxyapatite coating in Hanks solution, and
found that the corrosion rate of coated samples is more
than two times larger than that of uncoated ones.
However, in both studies, calcium phosphate coatings
were prepared through high-temperature treatment. The
effect of titanium oxide lm formed in the heat
treatment process on the corrosion resistance of
titanium materials was not considered. Sousa and
Barbosa [17] investigated the inuence of the addition
of calcium ions and phosphate anions into the solution
on the passive current of chemically passivated titanium
and found that calcium phosphate did not change the
passive current at potentials lower than 1.5 V, but did
ARTICLE IN PRESS
www.elsevier.com/locate/biomaterials
0142-9612/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biomaterials.2005.05.047

Corresponding author. Tel.: +902 585 1156; fax: +902 585 1114.
E-mail address: sharon.roscoe@acadiau.ca (S.G. Roscoe).
cause it to decrease between 1.5 and 3.0 V vs. SCE. The
electrochemical impedance spectroscopy (EIS) measure-
ments obtained at 1.0 V showed that calcium phosphate
signicantly decreased the charge transfer resistance
(R
ct
) of the prepassivated titanium electrode [17].
However, the inuence of this calcium phosphate
surface deposition on the corrosion resistance of the
titanium implant around the open circuit potential
(OCP) is still unknown.
Furthermore, titanium implants inserted into a human
body are usually surrounded with blood-rich tissue, and
the serum proteins in blood may also inuence the
corrosion of the implant materials. The inuence of serum
or serum proteins on the corrosion resistance of titanium
has been studied under different experimental conditions.
Different results (no inuence, increase or decrease) have
been reported [2,11,12,1720]. Clark and William [20]
investigated the corrosion of titanium powders in neutral
saline using atomic absorption spectrophotometry and
found no signicant inuence of serum proteins on the
corrosion of titanium. Okazaki et al. [11] investigated the
passivation behavior of several titanium alloys and found
no major variations between the passive current in saline
and in calf serum. Sousa and Barbosa [17] found serum
proteins brought no change to the anodic passivation
current of the calciumphosphate-coated titanium elec-
trode which had been pre-passivated in HNO
3
. However,
they also found that the presence of proteins signicantly
increased the R
ct
of the calcium phosphate-coated
titanium at 1.0 V. More research therefore needs to be
done to determine why the polarization results differ from
the impedance results. A number of investigations have
shown the corrosion resistance of titanium in a uoride
solution to be improved in the presence of albumin
[2123]. Contu et al. [2] investigated the inuence of fetal
bovine serum on the corrosion of titanium at OCP with
the EIS technique and reported that the addition of serum
increased the corrosion resistance at OCP by 23 times.
They attributed this result mainly to the ability of proteins
to bind free oxygen ions that move toward the metal/
oxide interface. Williams et al. [24] investigated the
corrosion resistance (R
p
) of titanium in 0.9% saline at
OCP using the linear polarization technique. They
reported that the presence of calf serum increased the
corrosion rate of the sanded commercially pure titanium,
which they attributed to the ability of some proteins to
form complexes with titanium ions. The similarity of the
electrodes and experimental conditions used in both
studies [2,24] suggests more research needs to be done to
clarify the contradictory results.
The acceleration effect of serum proteins on the anodic
dissolution of various metals has already been reported
[20,2527]. Clark and Williams [20] found that the
corrosion of cobalt and copper was greatly increased by
the presence of two serum proteins: serum albumin and
brinogen. Woodman et al. [27] reported that the
predominant corrosion products from 316 L stainless
steel were organometallic complexes with serum pro-
teins. Results obtained in our laboratory have shown
that the adsorption of serum albumin and b-lactoglobu-
lin onto implant-grade stainless steel caused an increased
corrosion rate, which was also attributed to the chelating
effect of proteins with metal ions [25,28,29].
Up to now, only a few reports [12,17] have been found
about the inuence of a calcium phosphate surface
deposit formed without high-temperature treatment on
the corrosion resistance and electrochemical polariza-
tion behavior of titanium around the OCP. Although
many studies have been made on the inuence of serum
on the corrosion resistance of titanium, and the effects
were usually attributed to the presence of serum
proteins, little work has been done with serum proteins
themselves, especially the inuence of serum proteins on
the cathodic and anodic polarization behavior of bare
titanium around the OCP.
The major goal of this study was to investigate the
inuence of calcium phosphate surface deposits and the
effect of the two major serum proteins, bovine serum
albumin (BSA) and brinogen, on the corrosion
resistance and electrochemical behavior of commercially
pure titanium in the potential range around the OCP. In
addition, the inuence of calcium phosphate deposits
and serum proteins on the anodic current of titanium
over a wide potential range, extending beyond the range
normally encountered for titanium in the body, was
used to obtain a comprehensive description of these
processes that may be encountered in industrial proces-
sing of biological material.
2. Experimental
2.1. Reagents and solutions
The stock solutions of BSA (Sigma Chemical Co. Product
no. A-0281), were prepared by dissolving an appropriate
amount of solid reagents in a phosphate buffer saline solution
(PBS), pH 7.4 (0.05 mol L
1
phosphate buffer+0.15 mol L
1
NaCl). The phosphate buffer was made by dissolving an-
hydrous monobasic potassium phosphate (Sigma Chemical Co.
P-5379) and NaCl (Fisher Scientic Company, S-271) in
conductivity water (Nanopure, resistivity 18.0 MOcm) and
adding 0.10 mol L
1
sodium hydroxide (made from5 N solution,
ACP Chemical Inc. Product No. S-3732) to adjust the pH of the
solution. A solution with 0.05 mol L
1
phosphate buffer (pH
7.4)+0.15 mol L
1
NaCl+2 mmol L
1
CaCl
2
(ACP Chemicals
Inc. C-0360) was used as an electrolyte to study the corrosion
behavior of titanium in the presence of calcium phosphate [17].
2.2. Electrochemical equipment
A single compartment electrochemical cell (volume ca.
120 mL) was used. The working electrode was a commercially
ARTICLE IN PRESS
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7351
pure titanium rod, (cp)-Ti (99.7%, 6.4 mm diameter, alpha
structure, JohnsonMatthey), sealed with epoxy resin
(Torr Seal, Varian Vacuum Products) to produce a two-
dimensional surface. The chemical composition of the
working electrode used in the research is given in Table 1.
The counter electrode was a large-area platinum electrode
(mesh) of high purity (99.99%, JohnsonMatthey), which was
degreased by reuxing in acetone, sealed in soft glass,
electrochemically cleaned by potential cycling in 0.5 mol L
1
sulfuric acid and stored in 98% sulfuric acid until use. A
saturated calomel electrode (SCE) was used as the reference
electrode, and all potentials in this paper are referred to
the SCE.
An EG&G PAR M273 potentiostat/galvanostat was used
for dc polarization measurements, and a Voltalab40 Dynamic
EIS and Voltammetry Electrochemical Laboratory (Radio-
meter) was used for EIS measurements.
2.3. Experimental methodology
All measurements were made in a PBS solution at 37 1C
without de-aerating. The level of dissolved oxygen was
assumed to be constant, since it was controlled by its partial
pressure in the atmosphere. The CaCl
2
solution or protein
solution was prepared in a separate container using a PBS
solution pH 7.4, and was allowed to equilibrate for at least
30 min in the constant temperature bath, tted with a Julabo P
temperature regulator, at the same temperature as the
electrochemical cell.
In order to obtain a reproducible surface, the electrode
was successively wet-polished down to 1500 gradation
using sand paper. After polishing, the electrode was thor-
oughly rinsed with Nanopure water, degreased in an ultra-
sonic bath containing ethanol for about 20 s, and then
again rinsed with Nanopure water. The electrode was then
immersed in a test solution at an OCP for 2 h in order to
reach a steady-state OCP value, followed by the EIS
measurement at the OCP and then the dc polarization
measurement in the cathodic direction. After the cathodic
polarization, the electrode was maintained at the open
circuit until the OCP value returned to that obtained
before the cathodic polarization. Polarization measure-
ments were then made in the anodic direction. After
the electrode was characterized by the electrochemical
techniques in the PBS solution, an aliquot of concentrated
CaCl
2
solution or protein solution was then added to the
electrochemical cell and the electrochemical measurements
were repeated.
All measurements were made in triplicate, with fresh
solutions and newly polished electrodes. The data is presented
with the reported average of the results with error bars
showing the spread in the triplicate measurements.
3. Results and discussion
3.1. Open circuit potential measurements
The dependence of the OCP on electrode immersion
time is shown in Fig. 1. Initially, the OCP rose rapidly
but then gradually slowed reaching a stable value.
Addition of BSA signicantly moved the OCP to
negative values, while the calcium phosphate deposit
and brinogen showed only a slight inuence on the
OCP.
The OCP values for titanium in PBS (Fig. 1) is
consistent with previously reported results obtained in
solutions without deaeration [30,31]. The passive oxide
lm of titanium has been investigated extensively
[6,8,3234]. When a freshly polished Ti surface is
exposed to moist air, a thin Ti-oxide lm is sponta-
neously formed on its surface. XPS analysis has shown
that the amorphous oxide lm, spontaneously formed
on Ti during polishing, consists of three layers: the rst
layer adjacent to metallic Ti is a TiO lm, on top of
which a Ti
2
O
3
lm is formed, and the third lm is TiO
2
which is in contact with the solution [35]. After the
titanium electrode was immersed into the electrolyte, a
titanium oxide lm began to grow on the electrode
surface. The protecting effect of the oxide lm, and
therefore the corrosion resistance of the electrode,
increased and eventually reached a relatively stable
value. The increase of corrosion resistance caused the
decrease of the anodic dissolution current of titanium.
As the OCP is determined by both the anodic and
cathodic reaction, according to the electro-neutrality
theory, the decrease of the anodic current will move
the OCP gradually in the positive direction so that the
cathodic current can be low enough to balance the
decreased anodic current. When the corrosion resistance
of the titanium oxide lm reached a relatively stable
ARTICLE IN PRESS
Table 1
Chemical composition of (cp)Ti electrode (mass%)
Material Ti C H Fe N O
(cp)Ti Bulk 0.10 0.015 0.5 0.05 0.40
Time / minutes
0 20 40 60 80 100 120
E

/

V

v
s
.

S
C
E
0.8
0.6
0.4
0.2
PBS
PBS + Ca ions
PBS + Fibrinogen
PBS + BSA
Fig. 1. Dependence of the OCP of Ti with time in: PBS;
PBS+2mmol L
1
CaCl
2
; PBS+0.05gL
1
brinogen; PBS+0.05gL
1
BSA.
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7352
value, the anodic current, and therefore the OCP, also
reached a stable value.
3.2. Electrochemical impedance spectroscopy
measurements
The EIS results at the OCP presented in Fig. 2a
showed only one time constant. Therefore, similar to
other references for titanium electrodes [2,29], the
Randles electrical-equivalent circuit (EEC) comprised
of only one time constant (Fig. 2b) was used to model
the experimental spectra, and good agreement between
experimental data and tted data was obtained.
Electronic elements in the gure have the following
meanings: R
e
is the resistance of the electrolyte between
the working and the reference electrode, R
p
is the
polarization resistance related to the rate of corrosion
reaction(s) at the OCP, and is inversely proportional to
the corrosion current, and CPE is the capacitance
represented by the constant-phase element (CPE).
Generally, the use of a CPE is required due to a
distribution of the relaxation times as a result of
inhomogeneities present on the microscopic level under
the oxide phase and at the oxideelectrolyte interface.
The R
p
and CPE obtained from the ts are shown in
Table 2. Addition of calcium signicantly increased the
R
p
by about 80%, and decreased the CPE. BSA and
brinogen brought no signicant change to the R
p
, but
increased the CPE. It is known that a calcium phosphate
layer will form on a titanium implant surface in neutral
solutions containing saturated calcium and phosphate
[13]. Elemental calcium and phosphorus were also
detected on the rinsed and dried electrode surface by
our EDX measurements. We therefore speculate that a
deposited calcium phosphate layer existed on the
electrode surface, which inhibited the electrochemical
reaction and increased the R
p
by covering the electrode
surface and blocking the mass transportation of oxygen
and/or reaction products to and/or from the electrode
surface.
In Table 2, the presence of calcium phosphate
decreased the CPE while in contrast the presence of
proteins increased the CPE value compared with that
obtained in the PBS solution alone. As described by
Contu et al. [2], the CPE in Fig. 2b contains the
contribution of both the capacitance of the oxide lm on
the titanium surface (C
f
) and the double layer capaci-
tance (C
dl
). These capacitances are in series giving the
equivalent capacitance as follows [2]:
CPE C
f

1
C
dl

1
.
When protein molecules adsorb onto the electrode
surface, a new capacitor (C
ad
) needs to be considered
in the equivalent circuit. If the surface is not homo-
geneously and compactly covered, the current will ow
through two parallel paths: the rst constituted by series
combination of C
f
and C
dl
, the second through the series
combination of C
f
and C
ad
[2]. As a result, the CPE
should be given by:
CPE C
f

1
C
dl

1
C
f

1
C
ad

1
.
In this case, the obtained CPE is possibly higher than
that in PBS because of the second half of the above
equation. However, if the coverage is homogenous and
compact, all three capacitors are in series and the CPE
will be given as:
CPE C
f

1
C
dl

1
C
ad

1
.
In this case, the CPE will be lower than that in PBS.
Since the proteins increased the CPE as shown in
Table 2, it is possible that the adsorbed proteins formed
a porous rather than a compact lm. On the other hand,
the calcium phosphate deposit may be a more compact
layer giving the lower CPE reported in Table 2.
3.3. dc Polarization measurements
Quasi-steady-state linear polarization measurements
were made to investigate the effect of the addition of
calcium ions on the electrochemical behavior of
ARTICLE IN PRESS
Z' / k
0 20 40 60 80 100 120 140
Z
'
'

/

k


140
120
100
80
60
40
20
0
PBS
PBS + Ca ions
PBS + BSA
PBS + Fibrinogen
R
p
R
e
CPE
(a)
(b)
Fig. 2. EIS plot of Ti electrode recorded at the OCP in: PBS;
PBS+2 mmol L
1
CaCl
2
; PBS+0.05 g L
1
brinogen; PBS+
0.05 g L
1
BSA.
Table 2
R
p
and CPE of (cp)Ti electrode in different solutions
PBS PBS+Ca ions PBS+BSA PBS+
Fibrinogen
R
p
(kOcm
2
) 201711 360715 18278 209712
CPE (mFcm
2
) 5574 3374 7776 7275
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7353
titanium around the OCP (Fig. 3). As the PBS
electrolyte used in this experiment is an oxygen-contain-
ing solution, the cathodic current recorded in the
presented potential region can be attributed to the
reduction of oxygen dissolved in the solution [36,37]. It
is seen that addition of calcium resulted in a signicant
current decrease in the cathodic range. Fig. 4 shows the
anodic polarization curves recorded in a PBS solution in
the absence and presence of calcium ions. Both curves
show similar behavior during the anodic polarization.
However, in the PBS solution, the anodic current rises
remarkably at about 2.6 V, while in the solution with
calcium, the anodic current begins to rise signicantly
only after about 3.1 V. Its current value is also lower
than that in PBS at potentials greater than 2.6 V.
Fig. 3 and Table 2 show that the calcium phosphate
deposit decreased the cathodic current and increased the
polarization resistance at the OCP. This can be
attributed to the covering and blocking effect of calcium
phosphate deposit on the cathodic reaction. Since the
deposited calcium phosphate decreased the cathodic
current, it might be expected to decrease the anodic
current as well through its blocking effect, causing a
decreased corrosion rate and increased polarization
resistance at the OCP as reported in Table 2.
Fig. 5 shows the dc polarization curves for titanium in
a PBS solution with and without addition of BSA or
brinogen. Both BSA and brinogen decreased the
cathodic current and moved the OCP in the negative
(cathodic) direction. The effect was most dramatic with
the addition of BSA (Fig. 5), which decreased the
cathodic current, increased the anodic current between
the OCP and the passive region, and caused a signicant
negative shift of the OCP. It is known that many organic
corrosion inhibitors prevent corrosion by forming an
adsorbed lm and blocking the mass transportation of
the corrosion process. It is also well known that proteins
have a high afnity for adsorption onto solid surfaces
[25,28,29]. The decreased cathodic current is therefore
interpreted as being due to adsorbed BSA molecules
covering the reaction sites and/or blocking the trans-
portation of dissolved oxygen to the electrode surface.
In order to examine the effect of protein concentra-
tion on the cathodic current, the current was recorded at
0.8 V after each incremental aliquot of BSA stock
solution (Fig. 6) and brinogen solution (Fig. 6, inset)
were made to the PBS solution. The potential of 0.8 V
was chosen since for all the polarization curves this
potential avoids the OCP and the diffusion control
potential regions. The current decreased with the
increase in protein concentration and reached a plateau
at 0.72 mM (0.05 g L
1
) for BSA and 0.037 mM
(0.01 g L
1
) for brinogen. The plateau threshold con-
centration for brinogen was much lower than for BSA,
i.e. brinogen required a lower equilibrium concentra-
tion in the bulk solution to reach a saturated adsorption
ARTICLE IN PRESS
E / V vs. SCE
1.0 0.8 0.6 0.4 0.2 0.0 0.2
l
o
g

(
j

/

A

c
m

2
)
7
6
5
PBS
PBS + Ca ions
Fig. 3. Polarization curves of Ti in PBS solutions with and without
2 mmol L
1
CaCl
2
recorded at a scan rate of 0.5 mVs
1
.
E / V vs. SCE
0 1 2 3 4
l
o
g

(
j

/

A

c
m

2
)
6
5
4
PBS
PBS + Ca ions
Fig. 4. Anodic polarization curves of Ti in PBS solutions with and
without 2 mmol L
1
CaCl
2
recorded at a scan rate of 1 mVs
1
.
E / V vs. SCE
1.0 0.8 0.6 0.4 0.2 0.0 0.2
l
o
g

(
j

/

A

c
m

2
)
7
6
5
PBS
PBS + Fibrinogen
PBS + BSA
Fig. 5. Polarization curves of Ti in: PBS; PBS+0.05 g L
1
brinogen;
PBS+0.05 g L
1
BSA recorded at a scan rate of 0.5 mVs
1
.
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7354
coverage, consistent with the relative size of the
molecules considering the projected areas for adsorption
(side-on orientation: BSA 46 nm
2
, brinogen 300 nm
2
).
In addition, the lower inhibiting effect on the cathodic
current of a brinogen-containing solution may be due
to a more open structure or lower degree of ordering of
adsorbed brinogen molecules at the surface compared
to BSA, which is a consequence of the difference
in protein structure. Fibrinogen is a brous molecule
with a long open structure, which fragments readily,
whereas BSA has an oval compact globular structure
that packs well and is fairly resistant to denaturation
during the adsorption process. Therefore a layer of BSA
is possibly more efcient in blocking the reaction sites
than the more open structure formed by the brinogen
layer.
If we suppose BSA has no inuence on the anodic
reaction of titanium, then, since BSA decreased the
cathodic current (i.e., reaction rate), the OCP should be
lower compared to that measured in the PBS solution
alone, and the corrosion resistance at the OCP should be
higher than that in PBS as we saw from Table 2 for the
solution with calcium ions. However, Table 2 shows that
BSA did not increase the polarization resistance of
titanium, which suggests that BSA may have increased
the anodic reaction around the OCP. It is known that
proteins can accelerate the dissolution of metals through
their chelation effects [20,2527], and therefore BSA
may also be able to accelerate the anodic dissolution of
titanium at the OCP by a similar chelating process. The
combination of the chelating effect on the anodic
reaction and blocking effect on the cathodic reaction,
results in a more negative OCP in the presence of BSA
than in PBS solution. Since titanium is a typical valve
metal, a lower OCP will cause a higher anodic current in
the potentiodynamic polarization curve between the
OCP and the passive range. Similar to the present
observed effect on the OCP by the adsorbed BSA,
Contu et al. [2] reported that the addition of serum
signicantly shifted the OCP of Ti in the negative
direction. This effect may be at least partially attributed
to the presence of BSA in the serum.
Similar to the effect observed with the presence of
calcium phosphate deposit in the high anodic range
(42.6 V in Fig. 4), addition of BSA and brinogen also
decreased the anodic current (Fig. 7) in the same high
anodic region. The sharp rise of the anodic current in
this potential range may result from the development of
fractures in the passive lm. The inuence of calcium
phosphate and proteins may be attributed to a blocking
effect on the crevices of the fractured passive lm.
4. Conclusion
Both calcium phosphate deposition and protein
adsorption decreased the cathodic current and the
anodic current in the high anodic potential range
(42.6 V). This inhibiting effect can be attributed to
their covering and blocking effect. Calcium phosphate
deposition increased the corrosion resistance of titanium
at the OCP. Although BSA dramatically shifted the
OCP in the negative direction, it did not signicantly
change the corrosion resistance at OCP. This is
attributed to the chelating effect by the protein on the
dissolution process and the inhibiting effect on the
cathodic reaction.
Acknowledgement
Grateful acknowledgment is made to the Nova Scotia
Health Research Foundation and the Natural Science
and Engineering Research Council of Canada for
support of this research.
ARTICLE IN PRESS
Concentration of BSA / mM
0.0 0.5 1.0 1.5 2.0 2.5 3.0
j

/

A

c
m

2
5
10
15
20
25
30
Concentration of Fibrinogen / mM
0.00 0.05 0.10 0.15 0.20
j

/

A

c
m

2
10
12
14
16
18
20
22
24
Fig. 6. Dependence of cathodic current at 0.8 V on the concentration
of BSA and brinogen (inset).
E / V vs. SCE
0 1 2 3 4
l
o
g

(
j

/

A

c
m

2
)
6
5
4
PBS
PBS + Fibrinogen
PBS + BSA
Fig. 7. Anodic polarization curves of Ti in: PBS; PBS+0.05 g L
1
brinogen; PBS+0.05 g L
1
BSA recorded at a scan rate of 1 mVs
1
.
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7355
References
[1] Rogers SD, Howie DW, Graves SE, Pearcy MJ, Haynes DR. In
vitro human monocyte response to wear particles of titanium
alloy containing vanadium or niobium. J Bone Joint Surg Br
1997;79B:3115.
[2] Contu F, Elsener B, Hohni H. Characterization of implant
materials in fetal bovine serum and sodium sulfate by electro-
chemical impedance spectroscopy. I. Mechanically polished
samples. J Biomed Mater Res 2002;62:41221.
[3] Gonzalez JEG, Mirza-Rosca JC. Study of the corrosion
behavior of titanium and some of its alloys for biomedical
and dental implant applications. J Electroanal Chem 1999;471:
10915.
[4] Krupa D, Baszkiewicz J, Sobczak JW, Bilinski A, Barcz A.
Modifying the properties of titanium surface with the aim of
improving its bioactivity and corrosion resistance. J Mater
Process Technol 2003;143:15863.
[5] Cai Z, Nakajima H, Woldu M, Berglund A, Bergman M, Okabe
T. In vitro corrosion resistance of titanium made using different
fabrication methods. Biomaterials 1999;20:18390.
[6] Cai Z, Shafer T, Watanabe I, Nunn ME, Okabe T. Electro-
chemical characterization of cast titanium alloys. Biomaterials
2003;24:2138.
[7] Grosgogeat B, Reclaru L, Lissac M, Dalard F. Measurement and
evaluation of galvanic corrosion between titanium/Ti6Al4V
implants and dental alloys by electrochemical techniques and
auger spectrometry. Biomaterials 1999;20:93341.
[8] Hanawa T, Hiromoto S, Asami K, Okuno O, Asaoka K. Surface
oxide lms on titanium alloys regenerated in Hanks solution.
Mater Trans 2002;43:30004.
[9] Khan MA, Williams RL, Williams DF. In-vitro corrosion and
wear of titanium alloys in the biological environment. Biomater-
ials 1996;17:211726.
[10] Ng BD, Annergren I, Soutar AM, Khor KA, Jarfors AEW.
Characterisation of a duplex TiO
2
/CaP coating on Ti6Al4V for
hard tissue replacement. Biomaterials 2005;26:108795.
[11] Okazaki Y, Tateishi T, Ito Y. Corrosion resistance of implant
alloys in pseudo physiological solution and role of alloying
elements in passive lms. Mater Trans, JIM 1997;38:7884.
[12] Lima J, Sousa SR, Ferreira A, Barbosa MA. Interactions between
calcium, phosphate, and albumin on the surface of titanium.
J Biomed Mater Res 2001;55:4553.
[13] Wen HB, Wijn JR, Cui FZ, Groot K. Preparation of calcium
phosphate coatings on titanium implant materials by simple
chemistry. J Biomed Mater Res 1998;41:22736.
[14] Metikos-Hukovic M, Tkalcec E, Kwokal A, Piljac J. An in vitro
study of Ti- and Ti-alloys coated with solgel derived hydro-
xyapatite coatings. Surf Coatings Tech 2003;165:4050.
[15] Lavos-Valereto C, Costa I, Wolynec S. The electrochemical
behavior of Ti6Al7Nb alloy with and without plasma-sprayed
hydroxyapatite coating in Hanks solution. J Biomed Mater Res
2002;63:66470.
[16] Fathi MH, Salehi M, Saatchi A, Mortazavi V, Moosavi SB. In
vitro corrosion behavior of bioceramic, metallic, and bioceramic-
metallic coated stainless steel dental implants. Dent Mater
2003;19:18898.
[17] Sousa SR, Barbosa MA. Corrosion resistance of titanium CP in
saline physiological solutions with calcium phosphate and
proteins. Clin Clin Mater 1993;14:28794.
[18] Khan MA, Williams RL, Williams DF. The corrosion behavior of
Ti6Al4V, Ti6Al7Nb and Ti13Nb13Zr in protein solutions.
Biomaterials 1999;20:6317.
[19] Khan MA, Williams RL, Williams DF. Conjoint corrosion and
wear of titanium alloys. Biomaterials 1999;20:76572.
[20] Clark GCF, Williams DF. The effects of proteins on metallic
corrosion. J Biomed Mater Res 1982;16:12534.
[21] Huang HH. Effect of uoride and albumin concentration on the
corrosion behavior of Ti6Al4V alloy. Biomaterials 2003;24:
27582.
[22] Ide K, Hattori M, Yoshinari M, Kawada E, Oda Y. The inuence
of albumin on corrosion resistance of titanium in uoride
solution. Dent Mater J 2003;22:35970.
[23] Takemoto S, Hattori M, Yoshinari M, Kawada E, Oda Y.
Corrosion behavior and surface characterization of titanium in
solution containing uoride and albumin. Biomaterials 2005;26:
82937.
[24] Williams RL, Brown SA, Merritt K. Electrochemical studies on
the inuence of proteins on the corrosion of implant alloys.
Biomaterials 1988;9:1816.
[25] Omanovic S, Roscoe SG. Electrochemical studies of the adsorp-
tion behavior of Bovine Serum Albumin on stainless steel.
Langmuir 1999;15:831521.
[26] Brown SA, Merritt K. Electrochemical corrosion in saline and
serum, Journal of Biomedical Materials Research. J Biomed
Mater Res 1980;14:1735.
[27] Woodman JL, Black J, Jimenez SA. Isolation of serum-protein
organometallic corrosion products from 316ss and HS-21 invitro
and invivo. J Biomed Mater Res 1984;18:99.
[28] Omanovic S, Roscoe SG. Interfacial behavior of beta-lactoglo-
bulin at a stainless steel surface: An electrochemical impedance
spectroscopy study. J Colloid Interf Sci 2000;227:45260.
[29] Jackson DR, Omanovic S, Roscoe SG. Electrochemical studies of
the adsorption behavior of serum proteins on titanium. Langmuir
2000;16:554957.
[30] Pan J, Thierry D, Leygraf C. Electrochemical and XPS studies of
titanium for biomaterial applications with respect to the effect of
hydrogen peroxide. J Biomed Mater Res 1994;28:11322.
[31] Aziz-Kerrzo M, Conroy KG, Fenelon AM, Farrell ST, Breslin
CB. Electrochemical studies on the stability and corrosion
resistance of titanium-based implant materials. Biomaterials
2001;22:15319.
[32] Hodgson AWE, Mueller Y, Forster D, Virtanen S. Electroche-
mical characterization of passive lms on Ti alloys under
simulated biological conditions. Electrochim Acta 2002;47:
191323.
[33] Marino CEB, Mascaro LH. EIS characterization of a Ti-dental
implant in articial saliva media: dissolution process of the oxide
barrier. J Electroanal Chem 2004;568:11520.
[34] Healy KE, Ducheyne P. Passive dissolution kinetics of titanium in
vitro. J Mater Sci-Mater Med 1993;4:11726.
[35] Pouilleau J, Devilliers D, Garrido F, Durand-Vidal S, Mahe E.
Structure and composition of passive titanium oxide lms. Mater
Sci Eng 1997;B47:23543.
[36] McMurray HN, Worsley DA, Wilson BP. Hydrogen evolution
and oxygen reduction at a titanium sonotrode. Chem Commun
1998;8:887.
[37] Hill MA, Butt DP, Lillard RS. The passivity and breakdown of
beryllium in aqueous solutions. J Electrochem Soc 1998;145:
2799806.
ARTICLE IN PRESS
X. Cheng, S.G. Roscoe / Biomaterials 26 (2005) 73507356 7356

Вам также может понравиться