Sie sind auf Seite 1von 9

Alkali and Heat Treatment of

Titanium Implant Material for Bioactivity


Ibrahim M. Hamouda, MSc, PhD1/Enas T. Enan, MSc2/
Essam E. Al-Wakeel, MSc, PhD3/Mostafa K. M. Yousef, MSc, PhD4
Purpose: This study was conducted to evaluate alkali- and heat-treated titanium implant material. Materials
and Methods: Ninety-eight square plates of commercially pure titanium were divided into three groups. Group
1 plates were left untreated, and groups 2 and 3 were subjected to anodization and alkali treatment for 24
and 48 hours, respectively. Treated specimens were then subdivided into three equal subgroups (a, b, and
c), which were heat treated for 1 hour at temperatures of 500C, 700C, and 800C, respectively. Changes
in the crystalline structure were analyzed using x-ray diffractometry. Surface roughness was measured
using a surface roughness tester. Selected specimens were immersed in a specially prepared simulated
body fluid for 10 days. Calcium and phosphorous deposition on the specimens was detected using energy
dispersive x-ray analysis. Results: Increasing the alkali treatment period and heat treatment temperature
positively affected surface roughness and formation of a bioactive sodium titanium oxide (sodium titanate)
layer on the titanium surface, especially after heat treatment at 800C. There was a significantly higher
calcium deposition on specimens of group 3 in comparison with those of groups 1 and 2. The results of pH
and ion concentration changes of the used simulated body fluid confirmed the results of energy dispersive
x-ray analysis. Conclusion: Alkali and heat treatment of titanium implant materials created better treatment
conditions for obtaining a bioactive implant material. Int J Oral Maxillofac Implants 2012;27:776784.
Key words: alkali treatment, bioactivity, heat treatment, implant materials, titanium

n recent years, titanium (Ti) has become a material


of great interest in dentistry and orthopedics. Ti has
many advantages, such as excellent biocompatibility,
corrosion resistance, and desirable physical and mechanical properties.1,2 The biocompatibility of Ti may
be attributed to its surface oxide film. This oxide film,
formed naturally in air, is a dense and stable anatase
(TiO2) with a thickness of few nanometers.3,4 Commercially pure Ti is available in four grades, which vary

1Professor

and Chairman, Department of Dental Biomaterials


and Restorative Dentistry, Faculty of Dentistry, Mansoura and
Umm Al Qura Universities, Mansoura, Egypt.
2D emonstrator, Department of Dental Biomaterials, Faculty of
Dentistry, Mansoura University, Mansoura, Egypt.
3 Professor, Department of Dental Biomaterials, Faculty of
Dentistry, Mansoura University, Mansoura, Egypt.
4Professor, Department of Metal Physics, Faculty of Science,
Mansoura University, Mansoura, Egypt.
Correspondence to: Dr Ibrahim M. Hamouda, Department
of Dental Biomaterials and Restorative Dentistry, Faculty of
Dentistry, Mansoura and Umm Al Qura Universities, Mansoura,
Egypt. Fax: +20-50-2260173. Email: imh100@hotmail.com

according to the oxygen (0.18% to 0.40% by weight)


and iron (0.2% to 0.5% by weight) content. These apparently slight differences in concentration have a
substantial effect on the physical and mechanical
properties.5
Among the most important applications of Ti and
its alloys is for dental implants, in addition to their use
for implant surface coatings, crowns, partial and complete dentures, and orthodontic wires. For successful implantation, the surface character of the implant
material becomes an important factor.5,6 In spite of its
excellent properties, Ti is usually bioinert, and integration between Ti and tissues is only a morphologic connection, although direct bone-implant contact, called
osseointegration, could occur.4
Considerable efforts have been directed toward
improving the strength of the bond between Ti implants and bone. Among these techniques is roughening of the Ti surface by coating, blasting with various
substances, acid etching, or combinations of these
treatments. As reported in many studies,79 strong interfacial bonding and active new bone formation have
been confirmed in the peripheral area around roughened implant surfaces. Other attempts to increase the

776 Volume 27, Number 4, 2012


2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

strength of the bone-implant bond have made use of


a Ti substrate that is either plasma sprayed or coated
with a thin layer of calcium phosphate ceramic as tricalcium phosphate or hydroxyapatite. The rationale for
coating an implant with calcium phosphate ceramic is
to produce a bioactive surface that promotes bone
growth and induces direct bonding between the implant and hard tissue, whereas the rationale behind
plasma spraying is to provide a roughened but biologically acceptable surface for bone ingrowth to ensure
anchorage in bone.10
It was recently claimed that Ti and its alloys can bond
to living bone by the formation of a bonelike apatite
layer on the surface of the metal without being coated
by hydroxyapatite but through chemical treatment
with sodium hydroxide (NaOH) solution, followed by
heat treatment.11,12 The hypothesis of this study was
to develop a new method that would increase bond
strength between Ti implants and surrounding bone.
Consequently, this study was designed to assess the
effects of alkali and heat treatments on Ti.

Materials and Methods


Specimen Preparation

Ninety-eight square plates of commercially pure Ti


(10 10 1 mm) were prepared by machining by
the manufacturer. They were polished with 400-grit
diamond paste and washed with pure acetone and
distilled water. They were classified into three main
groups: group 1 (14 specimens), as received (control);
group 2 (42 specimens), alkali treatment for 24 hours;
and group 3 (42 specimens), alkali treatment for 48
hours. Before alkali treatment, the specimens were
subjected to anodization to increase reactivity.13 The
specimens were chemically cleaned for 5 minutes in
5.5 mol/L of nitric acid with three drops of hydrofluoric acid, rinsed with distilled water, and dried at 40C.
They were immersed in an electrolyte solution (0.5%
by weight of hydrofluoric acid in water). A platinum
electrode (0.1 mm thick) was used as the cathode. A
direct current (20 V) was employed for 2 minutes at
room temperature.14

Alkali and Heat Treatments

A 5-mol/L concentration of NaOH aqueous solution


was prepared by dissolving 200 g of NaOH powder
in 1 L of distilled water. Each anodized specimen was
immersed in 15 mL of the prepared NaOH solution at
60C.15 Group 2 specimens were immersed in the prepared solution for 24 hours, while group 3 specimens
were immersed for 48 hours.11 To guarantee a uniform
and steady temperature of 60C during alkali treatment, a heating furnace (Stuart Scientific Furnace) was

Table 1Ion Concentrations and pH of SBF and


Human Blood Plasma16
Concentration (mmol/L)
Ion
Na+

SBF

Blood plasma

142.0

142.0

K+

5.0

5.0

Mg 2+

1.5

1.5

Ca2+

2.5

2.5

147.8

103.8

HCO3

4.2

27.0

HPO42

1.0

1.0

SO42

0.5

0.5

pH

7.40

7.207.40

Cl

used. After immersion, the specimens were washed


with distilled water and dried at room temperature.4
The alkali-treated specimens were divided into three
equal subgroups (14 specimens each): subgroup A was
heated at 500C, subgroup B was heated at 700C, and
subgroup C was heated at 800C, all for 1 hour. After
heat treatment, the specimens were allowed to cool
gradually to room temperature in the heat-treatment
furnace.15

Assessment of Crystal Structure

An x-ray diffractometer (XRD) (Heraeus T 5025) with a


copper target and nickel filter was used. The test was
conducted at an excitation voltage of 40 Kv and tube
current of 20 mA. Two specimens from each group
were subjected to irradiation, and the refracted and
transient rays were detected with the XRD. The individual peak positions (2), relative intensities (I/I),
and the corresponding interplanar spacing (d) were
obtained from a computer program printout. The crystalline structures of the alkali- and heat-treated specimens were identified and compared with those of the
control specimens.

Determination of Surface Roughness and


Morphology

A surface roughness tester (Surftest SJ-201P, Mitutoyo)


with a diamond stylus was used. Surface roughness
(Ra) was determined for each specimen at three different sites and the average was calculated. The cutoff
sampling length was set at 0.25 mm. The Ra value was
measured and calculated for 10 specimens from each
subgroup. Surface morphology was observed using a
scanning electron microscope (SEM) (JEOL JXA-840A)
at a magnification of 3,500.
The International Journal of Oral & Maxillofacial Implants 777

2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

650

650

600

600
T

500

500

450

450

400

350
300

350

200

200
150
T

50
T
0
10
20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80

300
250

150

400

250

100

RT

550

Counts/s

Counts/s

550

100

N
N

R
R R

RN

50
0
10
20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80

2 (stop)

2 (stop)

Fig 1 Representative x-ray spectrum for as-received Ti (T).

Fig 2Representative x-ray spectrum for Ti specimen (T) subjected to alkali treatment for 48 hours followed by heat treatment
at 800C. T = titanium; R = rutile; N = sodium titanium oxide.

Assessment of Surface Composition

ion concentrations in SBF were measured using an


atomic absorption spectrometer (Perkin-Elmer 2380).
One milliliter of each liquid sample of SBF was fed,
without dilution, into the apparatus, which combined
it with a stream of acetyleneair fuel and oxidant. The
mixture then passed into a burner in which the compounds making up the sample were broken into free
atoms. The absorbance values for calcium were 0.036,
0.072, and 0.278, while those for sodium were 0.217,
0.442, and 0.669 for 1, 2, and 3 ppm, respectively. To
measure the calcium ion concentration, a current of 10
mA and a wavelength of 422.7 nm were used. To measure the sodium ion concentration, a current of 8 mA
and a wavelength of 589 nm were used.

Five Ti specimens from each subgroup were soaked for


10 days in acellular simulated body fluid (SBF) with a
pH of 7.4. The composition of SBF and concentration of
the ions were nearly identical to those of human blood
plasma (Table 1).16 The SBF was prepared by dissolving 7.987 g of reagent-grade sodium chloride (NaCl),
0.352 g of NaHCO3, 0.622 g of potassium chloride
(KCl), 0.262 g of K2HPO43H2O, 0.17 g of MgCl26H2O,
0.278 g of calcium chloride (CaCl2), and 0.071 g of sodium sulfate (Na2SO4) into 1 L of distilled water and
buffered at a pH of 7.4 with tris(hydroxymethyl) aminomethane ([CH2OH]3CNH3) and hydrochloric acid at
36.5C.11 Each specimen was soaked in 25 mL of SBF
at 36.5C for 10 days in a Heraeus furnace. After soaking, the specimen was removed from the fluid, washed
with deionized water, and dried at room temperature.
The surface composition of the incubated specimens
was analyzed and compared with that of the control
specimens using energy dispersive x-ray (EDX) analysis
(INCA X-sight) in conjunction with the SEM.2,15 Calcium
and phosphorous (Ca-P) deposition on the treated Ti
specimens was assessed.

Evaluation of pH and Ion Concentration Changes

Five specimens from each subgroup were compared


with five specimens from the control group. Each specimen was immersed in 25 mL of SBF at 36.5C for 10
days and then removed from the fluid, which was then
collected and subjected to evaluation of the pH and
ion concentrations.15 Changes in pH were measured
using a pH meter (Consort p901). Calcium and sodium

Statistical Analysis

Data of Ra, pH, and calcium and sodium ion concentrations were collected. Means and standard deviations
were calculated for each group and compared by oneway analysis of variance (ANOVA) and least significant
difference (LSD) tests. Significance for all statistical
tests was set at = .05. Statistical analysis was performed with SPSS 14.0 (IBM) for Windows.

Results
XRD Findings

The XRD spectra obtained from as-received and from


alkali- and heat-treated Ti specimens are shown in Figs
1 and 2, respectively. These figures are plots of the relative intensity (counts per second) versus diffraction

778 Volume 27, Number 4, 2012


2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

Table 2Average d Spacing Values for Titanium


Oxide (Rutile) Lines Shown on the Diffraction
Patterns and the Corresponding ASTM Cards
Card/line

d ()

hkl

ASTM card
d ()

#44-1294

Table 3Average d Spacing Values for Sodium


Titanium Oxide Lines Shown on the Diffraction
Patterns and the Corresponding ASTM Card
(#11-0239)
Line

d ()

ASTM card
d ()

41

2.553

100

2.555

34

3.039

45

2.339

002

2.341

35

2.967

3.01
2.95

50

2.105

2.10

47

2.247

101

2.243

63

1.727

102

1.726

60

1.785

1.73

67

1.649

1.64

73

1.512

1.51

77

1.445

1.44

75

1.474

110

1.475

32

3.226

110

3.241

#87-0710
1
2

42

2.479

101

2.482

46

2.281

200

2.292

48

2.179

111

2.183

52

2.047

210

2.050

64

1.683

211

1.684

angle (2). Miller indices (hkl) of planes in the phases


present, which have been correlated with various diffraction peaks, are shown on the plots. The peak angles
show some variations in intensity of the diffracted
peaks because of preferred orientation. The XRD spectra for all specimens over the 2 interval from 20 to
80 showed the typical {100}, {002}, {101}, {102}, and
{110} peaks for the hexagonal Ti phase. Other peaks
{110}, {101}, {200}, {111}, {210}, {211}, {311}, {202}, and
{321}were attributed to the titanium oxide (rutile)
tetragonal phase. But other peaks, present at 2 of 34,
35, 50, 60, 67, 73, and 77 degrees, were attributed to
the sodium titanium oxide phase, as shown by d values, without indicating the Miller indices (hkl), because
no information was available about them, as indicated
by the ASTM card #11-0239. Tables 2 and 3 show the
calculated average d spacing values for each line corresponding to the lines reported on ASTM card #44-1294
for Ti, corresponding to rutile (titanium oxide) lines reported on ASTM card #87-0710, and corresponding to
sodium titanium oxide (sodium titanate) lines reported
on ASTM card #11-0239.

Surface Roughness Morphology

Mean surface Ra values and standard deviations for all


groups are shown in Table 4. The results indicated that
group 3C (alkali-treated for 48 hours and heat-treated
at 800C) had the highest mean Ra value, while group
1 (control) showed the lowest value. ANOVA showed
that there were significant differences (P < .0001)
among the surface roughness values of the studied
groups (Table 4). The LSD test showed that there were

Table 4Surface Roughness (Ra, Means SDs,


in m) of the Studied Groups
Group

Ra

LSD

0.174E 0.006

855.38

< .0001

0.013

2A

0.180E 0.010

2B

0.184E 0.011

2C

0.240D 0.007

3A

0.318C 0.008

3B

0.422B 0.015

3C

0.492A 0.008

Means with the same superscripts are not significantly different.


1 = control (as received); 2 = alkali treatment for 24 hours then heat
treatment at (A) 500C, (B) 700C, or (C) 800C; 3 = alkali treatment
for 48 hours then heat treatment at (A) 500C, (B) 700C, or (C) 800C.

no significant differences between groups 1 (control),


2A (alkali-treated for 24 hours and heat-treated at
500C), and 2B (alkali-treated for 24 hours and heattreated at 700C) (P > .05). There were significant differences between the control group and groups 2C
(alkali-treated for 24 hours and heat-treated at 800C),
3A (alkali treated for 48 hours and heat-treated at
500C), 3B (alkali-treated for 48 hours and heat-treated
at 700C), and 3C (alkali-treated for 48 hours and heattreated at 800C) (P .05). There were significant differences between group 2A and groups 2C, 3A, 3B, and 3C
(P .05). Significant differences were found between
group 2B and groups 2C, 3A, 3B, and 3C (P .05). Significant differences were also detected between group
2C and groups 3A, 3B, and 3C (P .05). There were also
The International Journal of Oral & Maxillofacial Implants 779

2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

2A

2B

2C

3A

3B

3C

Fig 3 SEM (3,500) of the studied specimens. 1 = control specimen; 2 = specimens subjected to alkali treatment for 24 hours followed by heat treatment at 500C (A), 700C (B), and 800C (C); 3 = specimens subjected to alkali treatment for 48 hours followed
by heat treatment at 500C (A), 700C (B), and 800C (C).

Ca

O
Ti

Ca

Ti

8
10 12
Energy (keV)

14

16

18

20

8
10 12
Energy (keV)

14

16

18

20

Fig 4 EDX spectrum for as-received Ti specimen (group 1).

Fig 5 EDX spectrum for Ti specimens subjected to alkali treatment for 24 hours followed by heat treatment at 800C.

significant differences between group 3A and groups


3B and 3C (P .05). There was a significant difference
between groups 3B and 3C (P .05).

Surface Composition

Surface Morphology

Figure 3 shows scanning electron micrographs of the


studied specimens. These images showed a difference
in surface texture between the control specimen and
the others. The control specimen (1) showed longitudinal elevations and depressions corresponding to the
direction of cutting, while images of specimens subjected to alkali treatment for 24 hours (2A, 2B, 2C) showed
some increase in surface irregularities and porosity. Ti
specimens that were subjected to alkali treatment for
48 hours (3A, 3B, 3C) showed a more granular texture.

Figures 4 to 6 are examples of EDX data showing peaks


corresponding to the different elements present in the
analyzed specimens after immersion in SBF. Weight
percentages for the detected elements are shown in
Table 5. Figure 4 shows that the control specimen was
composed mainly of Ti. Figure 5 shows EDX data for Ti
specimens subjected to alkali treatment for 24 hours
followed by heat treatment at 800C (group 2C). EDX
data for specimens of group 2C showed the highest
percentage of oxygen among the group 2 subgroups. In
addition, small percentages of calcium were detected
in subgroups 2B and 2C. Figure 6 shows EDX data for
specimens of group 3C (alkali-treated for 48 hours and
heat-treated at 800C). There were peaks correspond-

780 Volume 27, Number 4, 2012


2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

Table 5Average Composition of Ti Specimens

Elements (% by weight)
O
Ca
Ti

Group Titanium
Ca

Na P

8
10 12
Energy (keV)

14

16

18

20

Fig 6 EDX spectrum for Ti specimens subjected to alkali treatment for 48 hours followed by heat treatment at 800C (group
3C).

Table 6 pH (Means SDs) of SBF Used with


the Studied Groups
pH

LSD

Group

7.462E 0.008

1,812.41

< .0001

0.019

2A

7.954D

2B

7.952D

0.006
0.005

Oxygen

Calcium Phosphorous Sodium

98.54

2A

65.04

34.96

2B

65.73

33.98

0.20

2C

86.90

41.93

0.25

3A

83.67

39.36

0.30

3B

42.08

32.52

1.72

1.26

1.72

3C

58.75

89.13

5.56

0.4

1.27

1 = control (as received); 2 = alkali treatment for 24 hours then heat


treatment at (A) 500C, (B) 700C, or (C) 800C; 3 = alkali treatment
for 48 hours then heat treatment at (A) 500C, (B) 700C, or (C) 800C.

Table 7 Ca2+ and Na+ Concentrations (Means


SDs, in ppm) of SBF Used with the Studied
Groups
Group

Calcium ions (Ca2+)

Sodium ions (Na+)

5.226A 0.1

10.078D 0.001

2A

5.188A

0.2

10.080D 0.1

5.1 A,B

0.1

10.080D 0.002

2C

7.958D 0.008

2B

3A

7.980C 0.01

2C

5.026B 0.129

10.084D 0.1

3B

8.256B

3A

3.260C

0.2

14.340C 0.9

3B

2.620D

0.1

16.200B 0.3

3C

2.596D 0.1

17.814A 0.4

3C

0.03

8.326A 0.02

Means with the same superscripts are not significantly different.


1 = control (as received); 2 = alkali treatment for 24 hours then heat
treatment at (A) 500C, (B) 700C, or (C) 800C; 3 = alkali treatment
for 48 hours then heat treatment at (A) 500C, (B) 700C, or (C) 800C.

F
P
LSD

517.96

376.1

< .0001

< .0001

0.159

0.505

Means with the same superscripts are not significantly different.


1 = control (as received); 2 = alkali treatment for 24 hours then heat
treatment at (A) 500C, (B) 700C, or (C) 800C; 3 = alkali treatment
for 48 hours then heat treatment at (A) 500C, (B) 700C, or (C) 800C.

ing to calcium, phosphorous, and sodium, in addition


to titanium and oxygen. EDX data for specimens of
group 3C showed the highest percentage of oxygen
among the group 3 subgroups.

Changes in pH

Mean pH values and standard deviations for the SBF


for all groups are shown in Table 6. A comparison of
mean pH values of the tested specimens showed that
the SBF used with specimens of group 3C had the highest mean pH, while that used with specimens of group
1 (control) showed the lowest value. ANOVA showed
that there was a significant difference (P < .0001) between pH values of SBF used with the studied groups
(Table 6). The LSD test showed that there were no sig-

nificant differences between group 2A and groups 2B


and 2C (P > .05). No significant difference was detected
between groups 2B and 2C (P > .05). There were
significant differences between group 1 and groups
2A, 2B, and 2C (P .05). There were significant differences between group 1 and groups 3A, 3B, and 3C
(P .05). There were significant differences between
group 2A and groups 3A, 3B, and 3C (P .05). There
were significant differences between group 2B and
groups 3A, 3B, and 3C (P .05). In addition, significant
differences were found between group 2C and groups
3A, 3B, and 3C (P .05). Significant differences were detected between group 3A and groups 3B and 3C. There
was a significant difference between groups 3B and 3C
(P .05).
The International Journal of Oral & Maxillofacial Implants 781

2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

Calcium Ion Concentrations

Means and standard deviations of the calcium ion concentration in SBF used with all studied groups are shown
in Table 7. A comparison of mean calcium ion concentrations of the tested specimens showed that SBF used
with specimens of group 3C had the lowest mean concentration, while that used with specimens of group 1
showed the highest concentration. ANOVA showed that
there was a significant difference (P < .0001) between
calcium ion concentration values of SBF used with the
studied groups (Table 7). The LSD test showed that there
were no significant differences between groups 1, 2A,
and 2B (P > .05). No significant differences were found
between groups 2B and 2C (P > .05). No significant differences were seen between groups 3B and 3C (P > .05).
There were significant differences between group 1 and
groups 2C, 3A, 3B, and 3C (P .05). There were significant differences between group 2A and groups 2C, 3A,
3B, and 3C (P .05). There were significant differences
between group 2B and groups 3A, 3B, and 3C (P .05).
In addition, significant differences were found between
group 2C and groups 3A, 3B, and 3C (P .05). Significant
differences were found between group 3A and groups
3B and 3C (P .05).

Sodium Ion Concentration

Means and standard deviations of sodium ion concentrations in SBF used with all studied groups are shown
in Table 7. A comparison of mean sodium ion concentration values of the tested specimens showed that SBF
used with group 3C had the highest mean sodium ion
concentration, while that used with group 1 had the
lowest value. ANOVA showed that there was a significant difference between sodium ion concentrations of
SBF used with the studied groups (P < .0001). The LSD
test showed that there were no significant differences
between groups 1, 2A, 2B, and 2C (P > .05). There were
significant differences between group 1 and groups 3A,
3B, and 3C (P .05). There were significant differences
between group 2A and groups 3A, 3B, and 3C (P .05).
Significant differences were found between group 2B
and groups 3A, 3B, and 3C (P .05). There were significant differences between group 2C and groups 3A, 3B,
and 3C (P .05). In addition, there were significant differences between group 3A and groups 3B and 3C and
also between groups 3B and 3C (P .05).

Discussion
Biomedical and materials researchers have tried to
design the ideal surface to ensure long-lasting anchorage of implants. All bioactive materials developed up
to 1990 were based on calcium phosphate ceramics.16
It was later revealed that materials that form a calcium

phosphate layer, usually called a bonelike apatite, on


their surfaces in the living body bond to living bone
through this apatite layer, as it seems to activate bone
morphogenetic proteins and osteogenic cells to start
the cascade of events that result in bone formation.17
Apatite formation on a material can be induced by formation of functional groups such as TiOH, SiOH, TaOH,
and ZrOH on its surface. Based on these findings, bioactive Ti was prepared by forming sodium titanate,
which induces TiOH formation, on its surface via alkali
(NaOH) and heat treatments.16,18
XRD patterns for specimens of group 2 showed
sharp peaks with low intensity of the sodium titanate
phase, in addition to peaks of Ti and rutile, only after
heat treatment at 700C and 800C. Specimens that
were heat-treated at 500C did not show peaks of sodium titanate. On the other hand, the definition and
intensity of sodium titanate peaks were stronger in
specimens of group 3, which was treated with NaOH
for 48 hours. The broad peak of sodium titanate that
was shown after heat treatment at 500C may indicate
that this layer forms first in an amorphous form. This
form precipitated crystalline sodium titanate at 700C
and had fully crystallized at 800C, as indicated by the
sharper and more intense peaks present in the XRD
patterns, especially after 800C heat treatment. This indicated that 48 hours of NaOH immersion followed by
800C heat treatment produced the highest intensity
for the sodium titanate layer.
Leaching of Ti in NaOH results in the formation of a
hydrated titanium oxide gel layer containing alkali ions
on its surface (sodium titanate hydrogel layer).15,1921
This layer is dehydrated and condensed to form an
amorphous sodium titanate layer by heat treatment
below 600C.20,22 Regarding the time required for effective alkali treatment, one study11 showed disagreement with the present study. The authors concluded
that 24 hours of alkali treatment was sufficient for the
formation of a sufficient sodium titanate layer to start a
bioactive reaction on the Ti surface, while the results of
the present study revealed that 48 hours of alkali treatment was more effective for sodium titanate formation, as shown by the sharper and more intense sodium
titanate peaks in XRD patterns. This difference could be
attributed to variations among the experimental conditions, such as reactivity of the used solutions.
The hydrogel layer formed by alkali treatment is
mechanically unstable and requires further heat treatment to convert the gel layer into a more stable form.
The amorphous sodium titanate layer is converted
into crystalline sodium titanate and is rutile above
700C.15,1921 Sodium titanate results from the reaction
between titania (TiO2), which forms during anodization, and NaOH. When soaked in NaOH solution, titania react with OH, thus forming HTiO3; then titanate

782 Volume 27, Number 4, 2012


2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

hydroxide (HTiO3nH2O) is formed by the hydration of


HTiO3. These hydroxides are joined with sodium ions
in NaOH solution, and a porous network sodium titanate hydrogel layer is formed. After heat treatment,
a stable sodium titanate (Na2Ti5O11) layer is formed
eventually by the removal of nH2O from the sodium
titanate hydrogel layer.20,21
Surface roughness measurements and SEM in the
present study showed increased roughness, especially
in the case of specimens subjected to alkali treatment
for 48 hours. These results proved the presence of a
direct proportionality between the surface roughness
of Ti and the duration of alkali treatment, which can
be explained by the occurrence of a more prominent
reaction between Ti and NaOH with longer treatment
time.13 It was found that this increase in surface roughness favors implant fixation by inducing both boneanchoring and biomechanical stability.18,23
After soaking of the alkali- and heat-treated Ti
specimens in the prepared SBF, EDX performed in the
current study revealed the presence of calcium, phosphorous, and sodium on specimens that were subjected to alkali treatment for 48 hours, especially after
heat treatment at 700C and 800C. These results may
indicate the formation of a calcium phosphate layer
on the analyzed specimens. The presence of sodium
means that the reaction between sodium titanate and
the surrounding SBF was not yet completed, as more
sodium ions were still available for ion exchange and
formation of more calcium phosphate (apatite). On
the other hand, specimens that were alkali-treated for
24 hours showed only calcium deposition after 700C
and 800C heat treatment, while they did not show any
phosphorous peaks. These findings may indicate that
a longer alkali treatment period and subsequent heat
treatment above 700C led to a faster rate of apatite
formation on the treated specimens upon immersion
in SBF, indicating stronger bioactive behavior. In contrast to the treated groups, untreated specimens did
not show any calcium or phosphorous peaks, which
signals the absence of any bioactive reaction.
The formation of a calcium phosphate layer (bonelike apatite) on a materials surface is an essential requirement for bone growth on a synthetic material.
The role of this layer lies in the fact that it has a significant effect on cell adhesion and differentiation of
osteoblastlike cells, resulting in bone formation and a
tighter bone-implant bond.17,24 Regarding the effect of
the alkali treatment period on apatite formation on Ti,
the results of the present study were in disagreement
with a previous study,15 which detected apatite formation on Ti after alkali treatment for only 24 hours. On
the other hand, the present study revealed more obvious apatite deposition after 48 hours of alkali immersion. The different ability of alkali-treated Ti to induce

apatite nucleation could be explained by differences in


the Ti surface that depend on its treatment conditions.
The mechanism of apatite formation on alkali and
heat-treated Ti in SBF was interpreted in terms of an
electrostatic interaction between the Ti surface layer
and the ions in SBF. The previously formed sodium titanate layer releases sodium ions via exchange with
H3O+ ions in the SBF to form many Ti-OH groups on
the surface. As a result, the surface becomes negatively charged and reacts with the positively charged
calcium ions in the SBF to form calcium titanate. As calcium ions accumulate, the surface becomes positively
charged and reacts with the negatively charged phosphate ions to form amorphous calcium phosphate.
Because amorphous calcium phosphate is metastable
in SBF, it eventually transforms into stable crystalline
bonelike apatite.11,16
The pH and ion concentration analysis performed
for SBF used in the present research confirmed the
aforementioned mechanism of apatite formation. Results of pH analysis showed that immersion of group 2
specimens (alkali-treated for 24 hours) in the prepared
SBF caused an increase in its pH in comparison with
the results of the control group. SBF used with group 3
(alkali-treated for 48 hours) showed a greater increase
in pH, which showed its highest value on heat-treated
specimens at 800C. Elevated pH values of SBF after immersion of the treated specimens can be surely linked
to the ionic movements that took place between the
Ti surface and the surrounding solution, especially
the release of sodium ions, leading to changes in ionic
concentrations and therefore changes in pH.
The current data indicated that the ionic movement
in the soaking solution (SBF) leads to increase in its pH.
The alkali release and ion exchange in SBF resulted in
an increase in the pH of the surrounding fluid.15,23,24
This pH increase accelerates apatite nucleation by increasing the ionic activity product of apatite according
to the following equilibrium in SBF: 10Ca2+ + 6PO43+ +
2OH Ca10(PO4)6(OH)2.9
Measurement of calcium and sodium ion concentrations in the used SBF showed that specimens of
group 2 (alkali treated for 24 hours) did not change significantly in this regard. This may be attributed to the
presence of a weak ionic reaction among this group,
leading to weak bioactive behavior. Analysis of the results of group 3 (alkali-treated for 48 hours) showed
a notable decrease in calcium ion concentration simultaneous with a significant increase in sodium ion
concentration. A comparison of results of the different
heat treatment temperatures revealed that the highest
changes were related to Ti specimens that were heat
treated at 700C and 800C. Changes in calcium and
sodium ions concentrations may confirm the previously illustrated ion exchange cascade that ends with
The International Journal of Oral & Maxillofacial Implants 783

2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Hamouda et al

apatite deposition on treated Ti specimens by consuming calcium and phosphate ions from the SBF in which
they were soaked. Dependence of ionic changes on
the heat treatment temperature may be interpreted in
terms of more crystallization of the bioactive sodium
titanate layer at these temperatures (700C and 800C).
Based on the results of the present work, it can
be reported that ion exchanges between an implant
surface and SBF may strongly support the biochemical bonding theory of alkali- and heat-treated Ti with
the surrounding bonelike environment. In an in vivo
study,25 both treated and untreated porous Ti cylinders
were implanted in rabbit femoral condyles. Unexpectedly, there was no significant difference in bone ingrowth at the early postimplantation times of 2 and 4
weeks. Over time, however, the alkali- and heat-treated
implants showed increased osseointegration, whereas
the untreated implants did not. The authors attributed the delayed bone ingrowth, even with treated Ti,
to the type of bone in which implants were placed. It
was mentioned that the cancellous bone model used
in that study does not enhance bone ingrowth at early postimplantation periods, as it has low osteogenic
capacity, while the opposite occurs in cortical bone.
It was reported in the same study that osseointegration of untreated implants tends to be lost over time,
whereas the treated implants maintained osseointegration throughout the experiment.25

Conclusions
Based on the results and within the limitations of this
study, the following conclusions can be made.
1. Untreated titanium specimens showed the lowest
surface roughness values and no signs of any bioactive reaction.
2. Specimens treated in alkali for 48 hours showed
greater surface roughness and the formation of
a bioactive sodium titanate layer on the titanium
surface.
3. During heat treatment, the role of temperature
was obvious, particularly at 800C.
4. An increase in the length of the alkali treatment
period and an increase in the temperature of heat
treatment above 700C are recommended to increase the formation of a bioactive sodium titanate
layer.

References
1. Demirel F, Saygili G, Sahmali S. Corrosion susceptibility of titanium
covered by dental cements. J Oral Rehabil 2003;30:11621167.
2. Koik E, Fuji H. In vitro assessment of corrosive properties of titanium
as a biomaterial. J Oral Rehabil 2001;28:540548.
3. Carinci F, Volinia S, Pezzetti F, Francioso F, Tosi L, Piattelli A. Titanium-cell interaction: Analysis of gene expression profiling. J Biomed
Mater Res B Appl Biomater 2003;15:341346.
4. Feng B, Chen JY, Qi SK, He L, Zhao JZ, Zhang XD. Characterization
of surface oxide films on titanium and bioactivity. J Mater Sci Mater
Med 2002;13:457464.
5. Powers JM, Sakaguchi RL (eds). Craigs Restorative Dental Materials,
ed 12. Philadelphia: Mosby, 2006:5191, 97132, 327406.
6. Nishiguchi S, Fujibayashi S, Kim HM, Kokubo T, Nakamura T. Biology
of alkali-and heat- treated titanium implants. J Bone Joint Surg Br
2004;86:398402.
7. Yamagami A, Yoshihara Y, Suwa F. Mechanical and histologic
examination of titanium alloy material treated by sandblasting and
anodic oxidization. Int J Oral Maxillofac Implants 2005;20:4853.
8. Klokkevold PR, Nishimura RD, Adachi M, Caputo A. Osseointegration enhanced by chemical etching of the titanium surface. A
torque removal study in the rabbit. Clin Oral Implants Res 1997;8:
442447.
9. Orsini G, Assenza B, Scarano A, Piattelli M, Piattelli A. Surface
analysis of machined versus sandblasted and acid-etched titanium
implants. Int J Maxillofac Implants 2000;15:779784.
10. Anusavice KJ (ed). Phillips Science of Dental Materials, ed 11.
London: Saunders, 2003:563617, 759780.
11. Kokubo T, Kim HM, Kawashita M, Nakamura T. Bioactive metals:
Preparation and properties. J Mater Sci Mater Med 2004;15:99107.
12. Kim HM, Miyaji F, Kokubo T. Effect of heat treatment on apatiteforming ability of Ti metal induced by alkali treatment. J Mater Sci
Mater Med 1997;8:341347.
13. Wang G. Apatite-Forming Ability of Alkali-Treated Titanium Oxide
Coated Pure Titanium in Simulated Body Environment [thesis].
Kingston, Canada: Queens University, 2001.
14. Oh S-H, Finnes RR, Daraio C, Chen L-H, Jin S. Growth of nano-scale
hydroxyapatite using chemically treated titanium oxide nanotubes.
Biomater 2005;26:49384943.
15. Kim HM, Miyaji F, Kokubo T. Effect of heat treatment on apatiteforming ability of Ti metal induced by alkali treatment. J Mater Sci
Mater Med 1997;8:341347.
16. Kokubo T, Matsushita T, Takadama H, Kizuki T. Development of bioactive materials based on surface chemistry. J Eur Ceram Soc 2009;
29:12671274.
17. Vanzillotta PS, Sader MS, Bastos IN, Soares GA. Improvement of in
vitro titanium bioactivity by three different surface treatments.
Dent Mater 2006;22:275282.
18. Le Guehennec L, Soueidan A, Layrolle P, Amouriq Y. Surface treatments of titanium dental implant for rapid osseointegration. Dent
Mater 2007;23:844854.
19. Chosa N, Taira M, Saiton S, Sato N, Araki Y. Characterization of apatite formed on alkaline-heat-treated Ti. J Dent Res 2004;83:465469.
20. Lee BH, Kim YD, Lee KH. XPS study of bioactive graded layer in TiIn-Nb-Ta alloy prepared by alkali and heat treatments. Biomaterials
2003;24:22572266.
21. Xiao XF, Tian T, Liu RF, She HD. Influence of titania nanotube arrays
on biomimetic deposition of apatite on titanium by alkali treatment. Mater Chem Phys 2007;106:2732.
22. Jonasova L, Mller FA, Helebrant A, Strnad J, Greil P. Biomimetic
apatite formation on chemically treated titanium. Biomaterials
2004;25:11871194.
23. Geetha M, Singh AK, Asokamani R, Gogia AK. Ti based biomaterials,
the ultimate choice for orthopaedic implants: A review. Prog Mater
Sci 2009;54:397425.
24. Gil FJ, Padros A, Manero JM, Aparicio C, Nilsson M, Planell JA.
Growth of bioactive surfaces on titanium and its alloys for orthopaedic and dental implants. Mater Sci Eng 2002;22:5360.
25. Takemoto M, Fujibayashi S, Kokubo T, Nakamura T. Mechanical
properties and osteoconductivity of porous bioactive titanium.
Biomaterials 2005;26:60146023.

784 Volume 27, Number 4, 2012


2012 BY QUINTESSENCE PUBLISHING CO, INC. PRINTING OF THIS DOCUMENT IS RESTRICTED TO PERSONAL USE ONLY.
NO PART OF MAY BE REPRODUCED OR TRANSMITTED IN ANY FORM WITHOUT WRITTEN PERMISSION FROM THE PUBLISHER.

Das könnte Ihnen auch gefallen