Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Surface & Coatings Technology 231 (2013) 337–345

Contents lists available at ScienceDirect

Surface & Coatings Technology


journal homepage: www.elsevier.com/locate/surfcoat

Titanium surface modified by hydroxyapatite coating for dental implants


Kuo-Yung Hung a,⁎, Sung-Cheng Lo b, Chung-Sheng Shih a, Yung-Chin Yang c, Hui-Ping Feng a,⁎, Yi-Chih Lin a
a
Institute of Mechanical and Electrical Engineering, Ming Chi University of Technology, Taiwan
b
Department of Power Mechanical Eng., Ming Chi University of Technology, Taiwan
c
Dept. of Materials and Mineral Resources Eng., National Taipei U. of Technology, Taiwan

a r t i c l e i n f o a b s t r a c t

Available online 17 March 2012 The purpose of this study is to obtain a favorable combination of biocompatibility and mechanical properties
for dental implants by developing process parameters for plasma-sprayed hydroxyapatite (HA) coating on ti-
Keywords: tanium (Ti-6Al-4V ELI) surfaces. The plasma spraying experiments in this study, involving different process
Plasma spray parameters of HA coating (coating thickness up to approximately 120 μm), are divided into four stages. An
Hydroxyapatite coating immersion test for the HA coatings in a simulated body fluid (SBF) was performed after HA coating. After
Simulated body fluid
28 days, the crystallinity level of the HA coatings in the SBF increased from 54.88% to 74.39%, and the released
Dental implant
calcium ion concentration increased from 44.9 ppm up to 79.27 ppm. Phase III involved investigating the dif-
ferences in nozzle transverse speeds and surface speeds between disk- and rod (4.5 mm in diameter)-shaped
substrates for plasma HA spraying. The results of Phase III reveal that nozzle transverse speeds of 400 mm/s
and titanium surface speeds of 5 rpm may be optimal for the titanium rods. Finally, Phase IV entailed using
the parameters alternating from Phase III to perform HA plasma spraying on dental implants measuring
4.5 mm in diameter. An SEM morphology examination indicated that the coverage level of the HA coating
was nearly uniform and the thickness was approximately 47–130 μm. This study successfully applied plasma
spray technology to an HA spray for titanium surface modification. The evaluation and characterization of the
resulting HA coatings reveal that the Phase IV parameters may be used for HA-coating on titanium dental im-
plants. The coatings contained no significant phase components such as tricalcium phosphate (α-TCP, β-TCP)
or tetracalcium phosphate (TTCP).
© 2012 Elsevier B.V. All rights reserved.

1. Introduction (sandblasting with large grit and acid etching) [5–9] is currently the
main method. Hydroxyapatite (Ca10(PO4)6(OH)2, i.e., HA) is a calcium
Previous studies have explored the surface modification of titanium phosphate bio-ceramic material with nearly the same chemical compo-
dental implants for many years. Based on a study on titanium biocom- sition as human bone, and has remarkably high biocompatibility and
patibility conducted by Branemark in 1951 [1], titanium alloys have can be used to guide bone regeneration [10]. Accordingly, plasma-
become main materials for dental implants. Surface modification sprayed HA coating [4] is a promising method for enhancing the perfor-
can lead to favorable bone regeneration and integrity between the mance of dental implants.
bone tissue and implant surface. This technique not only retains the The power of plasma is determined according to the product of cur-
original biocompatibility of titanium alloys, but also improves the rent and voltage; therefore, the current is proportional to the power.
clinical performance of dental implants. Previous studies have recently The range of the current of a typical plasma-sprayed HA coating is
proposed numerous surface treatment methods, including sandblast- approximately between 350 and 1000 A [11]. Cizek et al. [12] demon-
ing, acid etching, electrochemical treatment, and thermal spray coating strated that high-power plasma spraying could elevate the temperature
[2–4]. Using the aforementioned surface modification techniques for of a flame and improve the degree of particle melting. Additionally, with
commercial dental implant products makes achieving most mechanical an accelerated flame and an increase of net power by 10 kW, the parti-
requirements and biocompatible properties possible. cle temperature can increase by 80 °C and the speed reaches 60 m/s.
Surface treatment is critical for dental implants. Among the implant Guessasma [13] indicated that the temperature of particles increased
surface treatment technologies in the dental implant market, SLA from 230 ± 272 °C to 263 ± 168 °C, and the speed of particles increased
from 221 ± 34 m/s to 324 ± 46 m/s as the current increased from 350 to
750 A. Therefore, as the current for plasma spray increases, both the
temperature and speed of particles increase accordingly.
⁎ Corresponding author at: Dept. of Mech. Engineering, 84 Gungjuan Rd., Taishan Dist.,
New Taipei City 24301, Taiwan. Tel.: +886 2 29089899x4514; fax: +886 2 29063269.
Quek [11], Tsui [14], Sun [15], and Yang [16] have investigated the
E-mail addresses: kuoyung@mail.mcut.edu.tw (K.-Y. Hung), effect of the power of plasma spray and current on HA coating. Tsui
hpfeng@mail.mcut.edu.tw (H.-P. Feng). [14] and Sun [15] have stated that increased power or current leads to

0257-8972/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2012.03.037
338 K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345

Fig. 1. Flowchart of the study procedure.

a decreased degree of crystallization on HA coatings. Tsui [14] also


proposed that an increase in the power of plasma spray reduces the
porosity and length of microfissures. However, Yang [16] claimed that
an increase of current in plasma spray enhances the HA degree of
crystallization.
Quek [11] indicated that the faster the lateral displacement of the
plasma flame is, the higher the porosity, the lower the degree of crystal-
lization, and the stronger the binding strength. Sun [15] investigated the
effectiveness of spraying distances of 8 to 16 cm, and determined that,
as the spraying distance increased both the degree of crystallization
and the hydroxyl content in HA reduced, resulting in an increase in
porosity and number of fractures. Kweh [17] demonstrated that an
increase in the spraying distance reduces the mechanical property of
the coating. The authors have found that the porosity and amount of
powder that have not melted have increased. In addition, an inhomoge-
neous deposition has formed at spraying distances between 12 and
14 cm, and that coatings with the optimal mechanical property have
been obtained at a spraying distance of 10 cm.
Taylor [18] revealed that certain materials, such as carbites, are not
Fig. 2. X-ray diffraction pattern of SULZER METCO powder and HA coatings on Ti-6Al-4V suitable for hydrogen gas coating, because hydrogen gas reacts with
ELI disks. coating materials. Guessasma [13] proposed that increasing gas flow
K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345 339

Table 2
Index of crystallinity (IOC) and coating thicknesses of hydroxyapatite coatings (HOC)
on Ti-6Al-4V ELI disks.

Exp XRD peak IOC Coating thickness (μm)

Powder 557
No. a1-HAC 262 47.0% 96.1
No. a2-HAC 320 57.5% 115
No. a3-HAC 400 71.8% 44.9
No. a4-HAC 289 51.9% 88.1
No. a5-HAC 328 58.9% 104
No. a6-HAC 480 86.2% 50.7

Therefore, using rotary fixtures resulted in a stronger bond for coatings


and substrates.
Process parameters for plasma-sprayed HA coating are intricate, and
significantly affect bonding strength, porosity, crystallinity, film thick-
ness, and surface roughness, as well as other factors. A literature review
revealed that while dental implants with plasma-sprayed HA coating
have been commercially available for many years, information about
the process parameters of screw-shaped implants has rarely been
reported in detail or completely [11–19]. Therefore, this study adopted
an experimental approach to investigate a suitable condition for HA
plasma spraying and to achieve proper characteristics of HA coatings
for dental implants. The experiments in this study included four stages
with various parameters of plasma-sprayed HA coating onto: disks in
Phase I and Phase II; cylindrical rods in Phase III; and screw-shaped
dental implants in Phase VI.

2. Materials and methods

This study presents a four-stage (Phases I, II, III, and VI) HA plasma
spraying experiment involving three substrate shapes (disks in Phases
I and II, cylindrical rods in Phase III, and screw-shaped dental implants
in Phase VI). Fig. 1 depicts the experimental process. This study used
the atmosphere plasma spray (APS) to coat the HA powder onto the
substrate. In addition, this study applied a rotary fixture to ensure
superior adhesion strength between the substrate and coating layer.

Fig. 3. FTIR spectra for HA powder from (a) SULZER METCO and (b) HA coatings for
2.1. Material samples
process No. b1 with immersion time in SBF.

Hydroxyapatite powder from SULZER METCO (XPT-D-703) was


while spraying can increase particle speed: as the gas flow increased used as a coating material in the plasma spray process for all substrates
from 30 liters per minute to 50 liters per minute, the average speed of in Phases I, II, III, and VI. Titanium (Ti-6Al-4V ELI, ASTM-F136) disks
the particles increased from 186 m/s to 269 m/s, and the temperature (25.4 mm in diameter and 3 mm in thickness, in Phases I and II ) and
of the particles increased from 2516 ± 131°C to 2526 ± 203°C. cylinders (25.4 mm in diameter and 55 mm in length, in Phases I and
Yang [19] investigated the effect of various relative motions on coat- II) for the bonding strength test conducted according to the standard
ings and used stationary (Z-shaped spraying path) and rotary fixtures test method ASTM C633, rods (4.5 mm in diameter and 15 mm in
(interlaced spraying path) for testing. The results demonstrated that length, in Phase III), and dental implants (4.5 mm in diameter, 15 mm
the surface coated using stationary fixtures was rougher, the degree of in length, in Phase VI) were all machined and subsequently grit-
crystallization was greater, the compressive residual stress was higher, blasted using angular corundum powders to fabricate adequate surface
the porosity was higher, and the distribution of pores was inhomoge- roughness and guarantee favorable mechanical adhesion between the
neous as compared with the surface coated using rotary fixtures. plasma-sprayed HA coating and the substrate. All of the grit-blasted
test specimens were cleaned in an acetone ultrasonic bath before
being sprayed with the HA powder.
Table 1
Process parameters of the atmospheric plasma sprayings on titanium (Ti-6Al-4 ELI)
disks (Phases I and II) and rods (Phase III).
2.2. Material characterization

Exp no. a1 a2 a3 a4 a5 a6 b1 c1 The adhesion strength of the plasma-sprayed HA coating was tested
Flow rate of primary gas (Ar) 50 50 50 50 50 50 50 50 according to the ASTM C633 standard, based on an average value from
(l/min) five measurements for each process condition. The degree of crystalliza-
Flow rate of secondary gas (H2) 14 6 0 14 6 0 14 14
tion (index of crystallinity, IOC) of plasma-sprayed HA coating (HAC)
(l/min)
Surface speed (Vs, rpm) 150 150 150 150 150 150 150 5 was calculated, based on X-ray diffraction analysis. The HA coating
Transverse speed (Vt, mm/s) 5 5 5 5 5 5 5 400 topography was performed using a scanning electron microscope
Flow rate of powder carrier gas (Ar) 3 3 3 3 3 3 3 3 (SEM: HITACHI S-2150). Chemical and crystallographic analysis was
(l/min) conducted using X-ray diffraction (XRD) (PHILIPS X'Pert Pro, Cu-Kα
Powder feed rate (rpm) 10 10 10 27 27 27 10 10
1.5418 Å, 40 kV, 30 mA), a Fourier transform infrared spectrometer
340 K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345

a d

b e

c f

6 µm

Fig. 4. Surface morphologies of HA coatings on Ti-6Al-4V ELI disks (4000×); process Nos.: (a) a1, (b) a2, (c) a3, (d) a4, (e) a5, and (f) a6.

(Perkin Elmer FTIR spectrophotometer), and an induction-coupled 3. Results and discussion


plasma atomic emission spectrometer (ICP-AES; Perkin Elmer Optima
2100DV). Three-dimensional surface roughness was tested using a 3D 3.1. Hydroxyapatite powder
surface analysis instrument (Sub-Nanometer 3D Optical Profiler Chro-
ma 7502) for the specimens of all plasma spraying conditions in this The sharp peaks of the HA phase in the XRD pattern of SULZER
work. Results are based on the average value Sa (the arithmetic mean METCO HA powder (Fig. 2) reveal that the main phase is purely HA.
between the roughness surface and its mean surface) and Sq (root Concerning the FTIR spectra for the HA powder, Fig. 3 provides informa-
mean square) from five measurements. tion on the molecular groups of the bands at approximately 567 and
The in vitro test in this study involved immersing the HA-coating 1089 cm− 1, which are associated with the vibration modes of the phos-
specimens in a simulated body fluid SIGMA H8264 (SBF; HBSS, Hank's phate group and approximately 630 and 3570 cm− 1 to hydroxyl group
balanced salt solution) for 30 ml/cm2 at 37 °C and 5% carbon dioxide in the HA powder. Fig. 3a shows the FTIR spectra for HA powder from
in a carbon dioxide incubator (CO2 Incubator; Thermo THERM FORMA SULZER METCO.
310) to observe the reaction response over different time periods (7,
14, 21, and 28 days). The purpose of the in vitro test was to inspect 3.2. Plasma-sprayed HA coating on disk-shaped substrates (Phase I)
the dissolving phenomenon caused by the amorphous phase of HA. A
favorable HA coating is required to achieve a high-quality biointegra- The six different conditions of plasma spraying in Phase 1 were
tion effect and to minimize the soluble phase. selected for HA coating onto titanium alloy (Ti-6Al-4V ELI) disk-shaped
K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345 341

a d

b e

c f

50 µm

Fig. 5. SEM cross-sectional images of HA coatings on Ti-6Al-4V ELI disks (4000×); process Nos.:(a) a1, (b) a2, (c) a3, (d) a4, (e) a5, and (f) a6.

substrates, based on our experience [19] and previous studies [11–19]. might create a lower crystallinity in the HA coating. A high temperature
These conditions appear as Exp No. a1–a6 in Table 1. The optimal process in plasma spraying because of a higher content of hydrogen can lead to a
parameters for the disk-shaped specimens were then evaluated, based on change in the phase composition of HA-sprayed coatings. This, in turn,
the resulting characteristics of the HA coatings regarding crystallinity, may form phases such as αTCP and βTCP. Hence, a higher powder feed-
surface roughness, morphology, thickness, and adhesion strength. ing rate may cause the HA powder to melt incompletely, thus leading to
a lower crystallinity level of HAC.
3.2.1. Crystallinity and purity
The index of crystallinity (IOC) value of HA coatings was derived by
3.2.2. Surface morphology
analyzing the peaks in the XRD pattern (Fig. 2). All of the HA coatings
The SEM surface morphologies of the HA coatings in Fig. 4 show that
were produced using six groups of process parameters with the same
the HA powder melted completely, forming splattered shapes for
HA phase peaks, but the crystallinity level for each coating was based
process Nos. a1 (Fig. 4a) and a4 (Fig. 4d) because of a higher content
on the differences in the peak height and broadening of XRD patterns.
of hydrogen in the processing. By contrast, the coatings produced
The calculation formula of the IOC is
using process Nos. a3 (Fig. 4c) and a6 (Fig. 4f) exhibited no splattered
morphology after plasma spraying with no hydrogen content. Thus,
IOC ¼ IHAC =IHAP  100% ð1Þ
the adding hydrogen or reducing the powder feed rate enhanced the
melting effect of HA powder because of an increase in heat transfer.
IHAC and IHAP are the XRD phase peaks of the HA coating layer and
This approach also attained a more favorable droplet spread, and
powder, respectively.
denser, more splatter-shaped HA coatings caused by powder particle
Table 2 reveals the crystallinity level and coating thickness of the
flattening.
hydroxyapatite coatings (HAC). In addition to the original HA phase in
the powder, some other phases, such as α-TCP and β-TCP, were intro-
duced into the HA coatings after plasma spray coating. These phases 3.2.3. Coating thickness
may be induced by a high temperature during plasma spraying. The The coating thickness of hydroxyapatite was measured using SEM
results of IOC of the hydroxyapatite coatings (HACs) meet the standard cross-sectional morphology (Fig. 5). Early resorption might occur
of ISO 13779–2:2000 [20], exceeding 45%. Experimental results also when the thickness is below approximately 75 μm, and a risk of delam-
reveal that higher hydrogen (H2) content during plasma spraying ination occurs above 200 μm. The average thickness values of the
342 K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345

a) c)
b) d)

e) f)

g) h)

i) j)

Fig. 6. (a–d) Concepts of the adhesion strengths measure mechanism between titanium alloy plates and HA films according to ASTM C633 method. (e–j) Fracture surfaces of
adhesion strength test on HA coatings; process Nos.: (e) a1, (f) a2, (g) a3, (h) a4, (i) a5, and (j) a6 (based on ASTM C633 specification; S: the side of HA-coated surface, L: the
other side of non HA-coated).

coatings ranged from 50.7 μm to 115 μm (Table 2), which is similar to


the range of 100 μm to 150 μm reported in relevant literature [21,22].

3.2.4. Adhesion strength between HA coating and titanium substrate


The purpose of adhesion strength tests is to understand the adhe-
sion effect of the HAC and substrate. Fig. 6 and Table 3 show the adhe-
sion strength levels between titanium alloy plates and HA films, as
measured according to the ASTM C633 method. The detailed proce-
dures are shown in Figs. 6a–d. The loading rod and substrate rod were
glued with a thermosetting adhesive (SULZER KLEBBI) after the sand-
blasting and coating procedures. The rods were then place into the
oven at 180 °C for 2 h. A universal tensile testing machine was used to
ensure the adhesion strength between the HA coating and the titanium
substrate.

Table 3
Adhesion strength test on HA coatings in Phase I (MPa).

Exp no. a1 a2 a3 a4 a5 a6

Average adhesion strength 37.4 34.9 39.9 20.2 33.2 44.8


Fig. 7. XRD patterns of the HA coatings for process No. b1 with immersion time in the SBF.
K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345 343

Table 4 Table 6
Index of crystallinity (IOC) and coating thicknesses (Phase II) with immersion time in Evolution of calcium ion concentration (Phase II) with immersion
SBF of hydroxyapatite coatings on Ti-6Al-4V ELI disks for the process No. b1. time in SBF.

Days in SBF IOC Coating thicknesses (μm) exp. Days in SBF HAC Ca-ICP (ppm)

Name No. b1 HAC No. b1-HAC

0 days 54.9% 128.2 0 44.90


7 days 56.1% 108.4 7 62.98
14 days 63.4% 104.7 14 71.38
21 days 53.7% 78.2 21 76.51
28 days 74.4% 75.6 28 79.27

The adhesion strengths of process Nos. a3 and a6 were higher than different periods (7, 14, 21, and 28 days). Peak analysis of the X-ray pat-
the others, but these may be not real values. In Fig. 7g, j can be explained terns reveals that the phase composition of all coatings in Phase II incor-
by the fact that the coating thickness was inadequate. A deficient thick- porated HA, α-TCP, and β-TCP. This may be due to a higher temperature
ness may be due to incomplete covering of the titanium alloy substrates generated by an increased spraying loop. Table 4 indicates the coating
by the HA coatings (Fig. 7g, j), causing direct contact somewhere crystallinity level for parameter No. b1 with immersion in the SBF.
between the titanium alloy plates and the bonding resin used for the According to ISO 13779–2:2000 [21], all coating crystallinity levels
bonding test. Therefore, the adhesion strengths are incorrect for process were appropriate for implants. The coating crystallinity of parameter
Nos. a3 and a6 with no hydrogen content during plasma spraying. The No. b1 was 54.88% for the HA coatings prior to immersion, and up to
results of Phase 1 show that the amount of hydrogen is a critical factor 74.39% after immersion in the SBF for a period of 28 days. This increase
in determining the melting degree of HA powder, and affects the valid- in the coating crystallinity after soaking in the SBF might be attributed
ity of the adhesion strength test. to the degradation of the non-crystalline part of the coatings.
The experimental results of plasma-sprayed HA coating on titanium The FTIR spectra results in Fig. 3b show that main absorption P–O
alloy disks in Phase I suggest that a higher hydrogen content or lower peaks decreased in the 1000–1100 cm− 1 and 550–600 cm− 1 regions
powder feeding rate allows the HA powder to melt completely. There- for the coatings immersed in the SBF for various periods. The minimal
fore, the surface morphology with splattered shaped areas is more obvi- P–O value occurred after soaking in the SBF for 7 days, and subsequently
ous, increasing the adhesion strength. The powder feed rate (27 rpm) increased because of the precipitation of the phosphate and deposition
for process Nos. a4, a5, and a6 was too high to melt the HA powder back to the substrate surface.
incompletely. Process Nos. a3 and a6 without hydrogen flow decreased
the plasma flame temperature and the powder melting rate. 3.3.2. Coating thickness
Process No. a1 produced HA coatings with an adhesion strength of Table 4 shows the thickness of the hydroxyapatite coatings, as
37.4 MPa, coating thickness of 96.1 μm, and crystallinity of 47%, where- measured according to the SEM cross-section images. The coating thick-
as process No. a2 achieved an adhesion strength of 34.9 MPa, coating ness without soaking in SBF was 128.16 μm, and decreased according to
thickness of 115 μm, and crystallinity of 57%. Thus, these coatings the number of days soaking in SBF.
meet the preliminary requirements for the properties of HA coatings
of dental implants and contain no significant components such as 3.3.3. Adhesion strength
α-TCP, β-TCP, or TTCP. The average adhesion strength level between titanium alloy
substrates and HA films using process No. b1 was 47.36 MPa, which is
3.3. Plasma-sprayed HA coatings on disk-shaped substrates (Phase II) higher than 37.42 MPa using process No. a1. The increase in the HA
coating adhesion strength using process No. b1 might be caused by
Process No. a1 of plasma spraying HA coating onto flat titanium alloy the spraying loop increase and poor heat conduction in titanium alloy
(Ti-6Al-4V ELI) substrates in Phase I achieved higher adhesion strength plates, which would elevate the substrate temperature and melt the
of 34.9 MPa. Therefore, Phase II involved using the process parameters powder completely.
of No. b1 presented in Table 1, adding spraying loops to ensure a coating
thickness of over 100 μm. The characterization of the HA coatings was 3.3.4. Surface roughness
performed to explore a fit process condition for HA coating onto cylin- Table 5 shows that the average value of the three-dimensional
drical rods in Phase III regarding surface morphology, crystallinity, surface roughness (Sa) of the HA coatings for process No. b1 without
purity, thickness, surface roughness, adhesion strength, and response soaking in SBF was 9.36 μm, and with soaking in SBF was 6.48 μm for
in the SBF. 7 days and 8.41 μm for 28 days. These results demonstrate that the
surface roughness of the HA coating remained nearly unchanged after
3.3.1. Crystallinity and impurity of HA coatings soaking in SBF for 28 days. The minimum Sa occurred after the 7th
Fig. 7 shows the X-ray diffraction patterns of the HA coatings pro- day of immersion and subsequently increased. This is likely due to the
duced by plasma spraying No. b1 while immersed in the SBF for precipitation of the phosphate and deposition back to the substrates.
The main purpose of released ion analysis for plasma-sprayed HA
coatings is to examine the HA coating response to the SBF, which
could affect the biocompatibility. The non-coated parts of the test pieces
Table 5
should be covered with glue to avoid direct contact with the SBF. Other-
Surface roughness (Phase II) as a function of immersion time in SBF.
wise, the non-coated parts on the back of the titanium plates directly
Exp. No. b1 HAC touch the SBF, preventing the specimens from releasing relevant ions.
Days (in SBF) Sa (μm) Sq (μm) After immersion in the SBF for a period, we conducted the detec-
0 9.36 11.67 tion of ICP-AES for the HA coatings. We found no released ions of
7 6.48 8.25 the titanium, vanadium, or aluminum. Table 6 indicates that released
14 7.95 10.41 calcium ions from the HA coatings without soaking in SBF was
21 8.23 10.52 44.9 ppm, and increased from 44.9 ppm with no soaking and up to
28 8.71 11.47
79.27 ppm after soaking for 28 days. It depicts that partial amorphous
344 K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345

Fig. 8. (a) Cross-section image of the dental implant, with regions of A (fine threads), B (coarse threads), and C (coarse threads). (b) Coating thicknesses of the dental implant in
Region A for process No. c1. (c) Coating thicknesses of the dental implant in Region B for process No. c1. (d) Coating thicknesses of the dental implant in Region C for process No. c1.

phase was coated to the substrate and dissolves quickly in body fluids 4. Conclusions
and thus, adversely affects bone formation. A stable HA coating is
required for good biointegration by minimizing the soluble phases. This study investigated the characteristics of plasma-sprayed HA
Fig. 3b shows the FTIR spectra for HA coatings for process No. b1 coating on 4.5 mm diameter dental screw-type implants by conducting
with immersion time in SBF. It depicts that all OH− and PO43 − still a sequence of experiments using different shapes of titanium sub-
exist after 28 days. strates. The experimental results of this study suggest the following
conclusions.
3.4. Plasma spray coating on cylindrical rods (Phase III)
1. Plasma HA spraying successfully modified the surfaces of 4.5 mm
The results of process No. b1 suggest that process No. c1 (Table 1) diameter dental implants. The operational parameters were a gas
should involve altering the rotation speed of the set table and swing (Ar) flow rate of 50 l/min, gas (H2) flow rate of 14 l/min, powder car-
speed of the spraying nozzle to avoid excessive heating of titanium rier gas (Ar) flow rate of 3 l/min, powder feed rate of 10 rpm, surface
substrates. Process No. c1 involved using plasma spraying on titanium speed 15 rpm, and transverse speed of 400 mm/s.
rods to determine the spraying parameters for screw-shaped dental 2. The dental implants in this study had the following characteristics:
implants. adhesion strength of approximately 41.44 MPa, coating thickness of
Comparing flat titanium alloy substrates with cylindrical rod sub- approximately 47–130 μm, degree of crystallinity of 54.88%, surface
strates shows that using process No. b1 with a lower nozzle swing roughness (Sa) of approximately 6.20 μm, and no significant compo-
speed increased the temperature of the cylindrical specimens because nents of α-TCP and β-TCP phases.
of a higher plasma flame temperature. Therefore, Phase III increased 3. Increasing the amount of hydrogen (6–14 l/min) or decreasing the
the nozzle swing speed to 400 mm/s to avoid overheating the implants. powder feeding rate (10 rpm) in plasma HA spraying allowed
Conducting SEM analysis of the cross-section images of the specimens the HA powders to melt completely. In addition to exhibiting higher
revealed that the thickness of the HA coating was 115.83 μm. adhesion strength, the resulting coatings were denser, more splatter-
shaped, and uniform.
3.5. Plasma spray coating on screw-shaped dental implants of (Phase IV) 4. Regarding the immersion of HA coatings in SBF, the degree of crystal-
linity increased to 74.9% after soaking for 28 days. The calcium ion
Fig. 8a presents cross sectional images of the dental implants pro- concentration increased with soaking duration (49.9 ppm for no
duced using process No. c1 in Phase IV and embedded in an organic soaking, and up to 79.27 ppm after soaking for 28 days), with no
resin. The regions of A (fine threads), B (coarse threads), and C titanium, aluminum, or vanadium ions released.
(coarse threads) for the dental implant in Fig. 8a exhibit different 5. This study successfully coated different HA thicknesses on dental im-
thicknesses of plasma-sprayed HA coating, ranging from 47 μm to plants using an atmosphere plasma spray, ranging from 47 μm to
130 μm. The A, B, and C parts of the implant (Fig. 8b–d) are almost en- 130 μm. The implants were almost entirely and uniformly covered
tirely and uniformly covered by the HA coating. These results indicate by the HA coating. These results indicate high biocompatibility for
high biocompatibility for bone regeneration and insulation against bone regeneration and insulation against the ion release of titanium
the ion release of titanium implants. implants.
K.-Y. Hung et al. / Surface & Coatings Technology 231 (2013) 337–345 345

Acknowledgement [7] D.L. Cochran, D. Buser, C.M. Bruggenkate, Clin. Oral Implants Res. 13 (2002) 144.
[8] M.M. Bornstein, A. Lussi, B. Schmid, U.C. Belser, D. Buser, Int. J. Oral Maxillofac.
Implants 18 (2003) 659.
We would like to thanks the Cheng Yang and Hui-Yun Bor (Mate- [9] M.M. Bornstein, B. Schmid, U.C. Belser, A. Lussi, D. Buser, Clin. Oral Implants Res.
rials & Electro-Optics Research Division, Chung-Shan Institute of Sci. 16 (2005) 631.
[10] A. Ravaglioli, A. Krajewski, Chapman & Hall Press, 1992, p. 44.
& Tech., Longtan) support the Plasma Spray equipment. [11] C.H. Quek, K.A. Khor, P. Cheang, J. Mater. Process. Technol. 89–90 (1999) 550.
[12] J. Cizek, K.A. Khor, Z. Prochazka, Mater. Sci. Eng., C 27 (2007) 340.
References [13] S. Guessasma, G. Montavon, C. Coddet, Surf. Coat. Technol. 192 (2005) 70.
[14] Y.C. Tsui, C. Doyle, T.W. Clyne, Biomaterials 19 (1998) 2015.
[1] P.I. Branemark, B.E. Breine, R. Adell, B.O. Hansson, J. Lindstrom, A. Ohlsson, Scand. [15] L. Sun, C.C. Berndt, C.P. Grey, Mater. Sci. Eng., A 360 (2003) 70.
J. Plast. Reconstr. Surg. 3 (1969) 81. [16] C.Y. Yang, B.C. Wang, E. Chang, J.D. Wu, J. Mater. Sci. - Mater. Med. 6 (1995) 249.
[2] S.J. Ferguson, N. Broggini, M. Wieland, M. de Wild, F. Rupp, J. Geis-Gerstorfer, D.L. [17] S.W.K. Kweh, K.A. Khor, P. Cheang, Biomaterials 21 (2000) 1223.
Cochran, D. Buser, J. Biomed. Mater. Res. Part A 24 (2006) 291. [18] M.P. Taylor, Ph. D. Thesis, University of Birmingham, Birmingham, 1994.
[3] B. Grosgogeat, L. Reclaru, M. Lissac, F. Dalard, Biomaterials 20 (1991) 933. [19] Y.C. Yang, Ph. D. Thesis, National Cheng-Kung University, 2003.
[4] R.B. Heimann, Surf. Coat. Technol. 201 (2006) 2012. [20] International Organization for Standards, BS ISO 13779–2, 2000.
[5] D. Li, S.J. Ferguson, T. Beutler, D.L. Cochran, C. Sittig, H.P. Hirt, D. Buser, J. Biomed. [21] C.Y. Yang, B.C. Wang, T.M. Lee, J. Biomed. Mater. Res. (1997) 36.
Mater. Res. 60 (2002) 325. [22] D.K. Groot, High Tech Ceramics, 225, Elsevier, Amsterdam, 1987, p. 147.
[6] M. Roccuzzo, M. Bunino, F. Prioglio, S.D. Bianchi, Clin. Oral Implants Res. 12 (2001) 572.

You might also like