You are on page 1of 10

Materials Science and Engineering C 48 (2015) 337–346

Contents lists available at ScienceDirect

Materials Science and Engineering C


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

Mixed zirconia calcium phosphate coatings for dental implants: Tailoring


coating stability and bioactivity potential☆
Karoline Pardun a, Laura Treccani a,⁎, Eike Volkmann a, Philipp Streckbein b, Christian Heiss c,d,
Giovanni Li Destri e, Giovanni Marletta e, Kurosch Rezwan a
a
University of Bremen, Advanced Ceramics, Am Biologischen Garten 2, 28359 Bremen, Germany
b
University Hospital, Justus-Liebig-University Giessen, Department of Cranio-Maxillo-Facial Surgery, Klinikstrasse 33, 35385 Giessen, Germany
c
University Hospital of Giessen-Marburg, Department of Trauma Surgery, Rudolf-Buchheim-Strasse 7, 35385 Giessen, Germany,
d
Laboratory of Experimental Surgery, Kerkraderstrasse 9, 35392 Giessen, Germany
e
Laboratory for Molecular Surfaces and Nanotechnology (LAMSUN), Department of Chemistry, University of Catania and CSGI, Viale A. Doria 6, 95125 Catania, Italy

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

Article history: Enhanced coating stability and adhesion are essential for long-term success of orthopedic and dental implants. In
Received 4 August 2014 this study, the effect of coating composition on mechanical, physico-chemical and biological properties of coated
Received in revised form 4 November 2014 zirconia specimens is investigated. Zirconia discs and dental screw implants are coated using the wet powder
Accepted 5 December 2014
spraying (WPS) technique. The coatings are obtained by mixing yttria-stabilized zirconia (TZ) and hydroxyapa-
Available online 9 December 2014
tite (HA) in various ratios while a pure HA coating served as reference material. Scanning electron microscopy
Keywords:
(SEM) and optical profilometer analysis confirm a similar coating morphology and roughness for all studied coat-
Calcium phosphate ings, whereas the coating stability can be tailored with composition and is probed by insertion and dissections
Zirconia experiments in bovine bone with coated zirconia screw implants. An increasing content of calcium phosphate
Coating stability (CP) resulted in a decrease of mechanical and chemical stability, while the bioactivity increased in simulated
Bioactivity body fluid (SBF). In vitro experiments with human osteoblast cells (HOB) revealed that the cells grew well on
Dissolution all samples but are affected by dissolution behavior of the studied coatings. This work demonstrates the overall
Cellular biocompatibility good mechanical strength, the excellent interfacial bonding and the bioactivity potential of coatings with higher
TZ contents, which provide a highly interesting coating for dental implants.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction bone ongrowth is achieved [5–7]. As a result, fibrous layer may form
around the implant and induce aseptic loosening [4,8]. Besides surface
Dental implants undergo several load-bearing conditions, such as roughness inorganic coatings on the material surface have been
chewing pressure, and thus require high mechanical reliability. They proposed in literature to overcome the instability and promote direct
must be able to withstand both, peak loads and millions of low level attachment to bone tissue [9,10]. Calcium phosphate (CP) based mate-
cyclic stresses occurring during lifetime while chewing [1]. This can be rials, such as hydroxyapatite (HA) and tricalcium phosphate (TCP) are
achieved by the use of metallic materials; the vast majority of dental im- considered to be bioactive and thus stimulate bone regeneration [11].
plants currently used are made of titanium or titanium alloy. Both mate- However, pure CP coatings exhibit a poor stability and provide a weak
rials have been established for years and are characterized by a high bonding strength to the substrate [12–15]. Therefore, one challenge is
degree of biocompatibility, favorable mechanical properties and low to control the adhesion strength of the coating together with a good bio-
corrosion rates [2]. Nevertheless, several reports about questionable activity but without affecting the bulk properties of the substrate, so
biostability have been stated [3,4]. that long-term stability can be achieved.
Ceramics, usually yttria-stabilized zirconia (TZ), receive increased To overcome these drawbacks of pure CP coatings several studies
attention as dental materials due to their low toxicity and beneficial me- used TZ reinforced HA coatings on titanium to combine enhanced me-
chanical properties like high fracture toughness and bending strength as chanical properties with an increased osseointegration. The addition
well as satisfying esthetic requirements. Despite of the biocompatible of TZ to HA and subsequent sintering provides an enhanced coating sta-
properties, TZ is known to be bioinert and thereby no stimulation of bility as well as better coating–substrate adhesion caused by improved
particle-substrate interactions when applied on zirconia substrates
☆ The authors declare no conflict of interest.
[16–18]. Sintering, a thermally driven process, gives the opportunity
⁎ Corresponding author at: Am Biologischen Garten 2, 28359 Bremen, Germany. for diffusion processes between the particles of coating and substrate,
E-mail address: treccani@uni-bremen.de (L. Treccani). which in turn may have a great influence on the adhesion [19].

http://dx.doi.org/10.1016/j.msec.2014.12.031
0928-4931/© 2014 Elsevier B.V. All rights reserved.
338 K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346

Both, bone-implant bond strength and coating-implant bond 2.2. Coating characterization
strength, are two important factors for successful and strong anchorage
of the implant to bone tissue [20]. Thus, a poor stability within the coat- The surface morphology of the coatings was analyzed using a scan-
ing and a weak bonding of the coating to the substrate could cause fail- ning electron microscopy (SEM, Camscan Series 2, Obducat CamScan
ure or delamination due to the occurring forces during insertion and Ltd., Cambridgeshire, United Kingdom) at 20 kV operating in secondary
after implantation [21,22]. Therefore, coating composition and the coat- electron mode and previously sputter coated with gold for 60 s (K550,
ing characteristics involved are of great interest to achieve a high effec- Emitech, West Sussex, UK).
tiveness and long-term stability of the implant. Measurements of the surface roughness were characterized with an
Regarding the significance of coating stability and adhesion the aim optical profilometer (Plμ2300, Sensofar, Terassa, Spain). Areas of 477
of this study was to assess the effect of coating composition on mechan- × 636 μm2 were scanned and the average surface roughness was calcu-
ical, physico-chemical and biological properties of coated zirconia spec- lated according to ISO 25178.
imens. Different mixed coatings consisting of TZ and CP were prepared Crystal phase analysis was carried out by grazing incidence X-ray
by wet powder spraying (WPS) and subsequently investigated for mi- diffraction (GI-XRD) using an Ultima IV type III diffractometer (Rigaku,
crostructure, phase analysis, coating adhesion and mechanical strength. Tokyo, Japan) equipped with Cross Beam Optics (CBO) and Cu Kα radi-
In vitro dissolution behavior, and surface morphology in simulated ation. Each run was performed with 2 theta (2θ) values between 20°
body fluid (SBF) as well as proliferation and differentiation of human and 70° carried out in parallel beam mode with a fixed incident angle
osteoblasts (HOB) were examined to understand and evaluate the me- of 0.5°.
chanical and chemical stability, and bioactivity of different coating
compositions. 2.3. Characterization of mechanical properties

Coating adhesion was determined with scratch tests performed by a


2. Materials and methods pencil hardness tester (PH-5800, BYK-Gardner GmbH, Geretsried,
Germany) according to ISO 15184. The mechanical adhesion between
2.1. Coating preparation substrate and coating was qualitatively examined with a pencil (Vickers
hardness 90.6 ± 2.5 HV 0.2) and a pencil-shaped bovine femur (Vickers
The details of substrate and suspension preparation, as well as the hardness 89.8 ± 5.2 HV 0.2), which were fixed in a sclerometer and
coating process are described in Pardun et al. [18]. moved over the surface with a constant load of 7.5 N and a velocity of
Briefly, zirconia discs (diameter (Ø) = 15 mm, height (h) = 1 mm/s along a 12 mm track. Prior to SEM investigation graphite and
1.7 mm) were fabricated using commercially available TZ-3YSB-E pow- bone leftovers on the sample surface were burned out at 1000 °C and
der (Tosoh, Tokyo, Japan) and uniaxially pressed at 38 MPa. After press- 1400 °C, respectively.
ing, the substrates were isostatically densified at 1200 bar for 5 min and Biaxial flexural strengths of the coated specimens (Ø = 15 mm, h =
pre-sintered at 1100 °C for 2 h. Dental implants (length (l) = 17 mm, 1.7 mm) were tested with the ball on three balls (B3B) test using a uni-
Ø = 4.5 mm) prepared from a zirconia feedstock (INMAFEED K1012, versal testing machine (Zwick/Roell Z005, Ulm, Germany) with a con-
INMATEC Technologies GmbH, Rheinbach, Germany) were fabricated stant speed of 0.5 mm/min and metal balls of 4 mm radius. The
via injection molding and subsequently pre-sintered at 950 °C for 2 h flexural strength was calculated applying the following equation [23]:
(Fraunhofer IFAM, Bremen, Germany).
The different mixed suspensions for coating were prepared of HA 3  F  ð1 þ νÞ
σ max ¼
"4  π  t  
2
(Sigma-Aldrich Chemie GmbH, Munich, Germany) and TZ-3YS-E pow- !   #
der (TZ, Tosoh, Tokyo, Japan), coating compositions are shown in Ra 1−ν ðt=3Þ2 Ra 2
 1 þ 2 ln þ  1−  ð1Þ
Table 1. The aqueous ceramic suspensions consisted of different amounts t=3 1þν 2  Ra R
of HA and TZ and were stabilized with polyacrylic acid-based dispersant
(PAA, 12 mg/g ceramic powder, Syntran® 8220, Interpolymer GmbH, where: F is the applied load, ν the Poisson's ratio of the substrate mate-
Hassloch, Germany) and at pH 10 by the addition of ammonium hydrox- rial (here 0.32 for TZ), t the sample thickness, Ra the support radius and
ide solution (25% NH3 basis, Sigma-Aldrich). As reference material a R the sample radius. The strength was tested on 30 samples for each
pure HA coating was used. All suspensions were homogenized by coating. Furthermore, fracture surfaces of the B3B specimens were ob-
ultrasonication (Sonifier 450, Branson, Dietzenbach, Germany) for served under SEM to analyze the coating–substrate interface.
10 min (mixed coatings) and 3 min (pure HA coating).
The coating of pre-sintered substrates was done with WPS. The sub- 2.4. Insertion and dissection of coated zirconia dental implants in bovine rip
strates were fixed and coated using a double-action airbrush spray gun bone
(BD 183-K, Artistic Life, Boenen, Germany) with a working distance of
200 mm. Coating deposition was performed with an air pressure of The surface coated zirconia screws (n = 10) were inserted into
2 bar, a relative humidity of ~ 60% and an airbrush nozzle with Ø predrilled holes of five fresh bovine rip bones. The insertion was carried
0.8 mm. After coating the samples were dried for 24 h at room temper- out with drilling and insertion tools from the similarly macrodesigned
ature and afterwards sintered at 1500 °C for 2 h. All subsequent charac- implant system (BEGO Semados® RI; BEGO Implant Systems, Bremen,
terizations were performed on sintered samples. Germany) and performed according to the clinically approved protocols
provided by the manufacturer.
After careful dissection of the implants SEM was used to assess coat-
ing morphology and stability. To further examine the coating stability at
Table 1
Coating composition of the different mixed coatings. the thread tips, the same samples were embedded in resin (EpoFix,
Struers GmbH, Willich, Germany), grounded to the middle of the im-
Sample name TZ HA
plant and examined by SEM.
wt.% vol.% wt.% vol.%

CP – – 26.9 12.0 2.5. Dissolution behavior


TZCP 1:2 12.9 3.0 26.9 12
TZCP 1:1 21.1 5.1 21.1 9.9 To test the dissolution behavior of the various coatings, specimens
TZCP 2:1 29.5 7.5 15.4 7.5
were placed into sterile polystyrene culturing dishes (24-well
K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346 339

multidish, Nunc, Wiesbaden, Germany) and immersed in 1 ml Tris–HCl Mannheim, Germany), which is enzymatically cleaved to formazan
buffer up to 21 days. Tris–HCl buffer solution was prepared by dissolv- only by living cells. The WST-1 reagent was added to each well after me-
ing 13.25 g of Tris(hydroxymethyl)aminomethane (Sigma-Aldrich) in dium change (n = 6 for each kind of specimen and day) and incubated
500 mL double deionized water and buffered at pH 7.4 with 1 M hydro- at 37 °C and 9.3% CO2 for 2.5 h. At days 1, 4, 7 and 9 of culture the amount
chloric acid (HCl, Sigma-Aldrich) at 37 °C according to ISO 10993-14. of formazan was measured at 450 nm using a microplate reader
The samples were placed in an incubator at 37 °C (Inkubator 1000, (Chameleon, HIDEX, Turku, Finland).
Heidolph, Schwabach, Germany) with an integrated shaker (160 rpm, The investigation of cell number, alkaline phosphatase activity (ALP)
Unimax 1010, Heidolph). The supernatant was replaced by fresh Tris– and protein concentration was performed with cell lysates (n = 5 for
HCl solution at 12 different sampling points. The release of calcium each kind of specimen and day) after 4, 7 and 9 days of culture. Cell
ions (Ca2 +) into the supernatant was determined photometrically lysis was achieved with 1% Triton X-100 (Sigma-Aldrich) in 0.9% NaCl
using the ortho-cresolphthalein complexone method (Fluitest Ca CPC, (Sigma-Aldrich) and incubation for 1 h on ice.
Analyticon Biotechnologies AG, Lichtenfels, Germany) in accordance to The cell number on the ceramic samples was assessed by DNA anal-
the manufacturer's instruction. Ortho-cresolphthalein complexone ysis using the Quant-iT™ PicoGreen® dsDNA Kit (Life Technologies). An
forms a purple-colored complex in alkaline solution. The color intensity aliquot of the cell lysate was diluted with 1× TE buffer (10 mM Tris–HCl,
is proportional to the Ca2+ concentration and was measured at 578 nm 1 mM EDTA, pH 7.5) prior to addition of PicoGreen working solution.
(Multiscan GO, Thermo Scientific, Schwerte, Germany). Experiments Sample fluorescence was measured after an incubation of 5 min with
were carried out in replicates (n = 4). an excitation wavelength of 485 nm and an emission wavelength of
535 nm using a plate reader (Chameleon, HIDEX). The DNA concentra-
2.6. Immersion in simulated body fluid tion was correlated with the cell number using a calibration curve with
defined cell number.
Conventional simulated body fluid (cSBF) was prepared by dissolv- For the determination of the ALP activity an aliquot of the cell
ing the following salts NaCl, NaHCO3, KCl, K2HPO4·3H2O, MgCl2·6H2O, lysate was incubated with ALP substrate buffer, containing 6 mM
CaCl2 and Na2SO4 (all salts were purchased from Sigma-Aldrich). cSBF paranitrophenyl phosphate in 0.1 M glycine, 1 mM MgCl2, 0.1 mM
was prepared according to Oyane et al. and ISO 23317. The solution ZnCl2 at 37 °C for 30 min. The reaction was stopped with 1 M NaOH
was buffered at pH 7.4 with Tris–HCl. Each sample was immersed in and released 4-nitrophenol was determined photometrically at
40 mL of cSBF solution and incubated at 37 °C under static conditions 405 nm. A standard curve was generated with different concentrations
in a climate chamber (KBF 115, Binder, Tuttlingen, Germany). After im- of 4-nitrophenol. All chemicals for the ALP activity were purchased from
mersion for 4, 7, 10, 14 and 21 days the samples were gently rinsed with Sigma-Aldrich. The protein contents were assessed with the Pierce®
double deionized water and dried at room temperature. The newly BCA Protein Assay Kit (Thermo Scientific). An aliquot of the cell lysate
formed mineral layer was observed by SEM (Carl Zeiss AG, Oberkochen, was incubated with working reagent at 37 °C for 30 min and afterwards
Germany) and the chemical composition was determined by energy- measured at a wavelength of 570 nm. The specific ALP activity was cal-
dispersive X-ray spectroscopy (EDX, INCA PentalFETx3, Oxford Instru- culated by normalization to the protein content.
ments, Tubney Woods, UK). In addition, the supernatant was collected
after each sampling day and Ca2+ concentration was measured photo-
2.8. Statistical analysis
metrically using Fluitest Ca CPC (Analyticon), as discussed in Section 2.5.
All collected data are expressed as mean values and standard devia-
2.7. In vitro biological investigation
tion. Statistical analyses were performed using one-way ANOVA follow-
ed by Tukey's test (Minitab 16, Minitab Inc., Pennsylvania, USA) for
In vitro investigations with human osteoblast cells (HOB, Provitro
multiple comparisons between the different coated groups. A confi-
GmbH, Berlin, Germany) were conducted up to 9 days to assess the bio-
dence level of 5% was considered to indicate a statistically significant
compatibility of the different coated samples. All samples were heat
difference.
sterilized at 200 °C for 2 h prior cell seeding. Thermanox® (∅ =
15 mm, surface roughness 0.03 ± 0.0 μm, Thermo Fisher Scientific
Inc., Bonn, Germany) was used as reference material. 3. Results
HOBs were grown in Dulbecco's modified Eagle's minimal essential
medium (D-MEM, Gibco, Life Technologies GmbH, Darmstadt, 3.1. Coating characterization
Germany) containing 10% heat-inactivated fetal calf serum (FCS,
Sigma-Aldrich) and 1% antibiotic–antimycotic (Gibco) in an incubator Surface morphology of all studied coatings was analyzed by SEM to
(C200, Labotect Labor-Technik-Göttingen GmbH, Göttingen, Germany) understand the influence of TZ:CP ratio on the resulting microstructure
at 37 °C, 9.3% CO2 and 95% relative humidity [24]. The culture medium (Fig. 1). SEM micrographs of all sample surfaces revealed an undulated,
was changed every two days. Cells were seeded onto the different ce- rough and porous morphology. CP, TZCP 1:2, TZCP 1:1 and TZCP 2:1
ramic samples at a density of 2 × 104 cells/ml in 24-well multidishes. showed an average surface roughness of 9.0 ± 1.1 μm, 7.4 ± 0.8 μm,
To evaluate the cell morphology and growth immunofluorescence 6.4 ± 0.5 μm, and 5.8 ± 0.4 μm, respectively. Furthermore, pores sizes
staining was carried out at days 1, 4 and 7 of culture. Samples (n = 3) between 2 and 20 μm were observed. Formation of cracks along the
were rinsed with 37 °C PBS twice and fixed with 4% paraformaldehyde coating surface was also detected on samples CP and TZCP 1:2; the
(PFA) (Riedel-de Haën, Seelze, Germany) for 15 min at room tempera- width of the cracks was less than 0.5 μm. The thickness of the coating
ture. After washing with PBS, the cells were permeabilized for 3 min varied between 33 and 55 μm.
with 0.5% Triton X-100 (Sigma-Aldrich). Cell nuclei and cytoskeletal GI-XRD analyses performed after sintering at 1500 °C are given in
actin were stained with 4,6-diamidino-2-phenylindole (DAPI, 1:2500, Fig. 2. The bottom diffractogram exhibits XRD pattern of the pure CP
Sigma-Aldrich) and Alexa Fluor® 488-Phalloidin (1:100, Life Technolo- coating where all peaks are consistent with the presence of β-TCP and
gies), respectively, for 45 min at room temperature in the dark. After ad- α-TCP. TZCP 1:2 consisted of cubic calcium zirconium oxide and β-
ditional washing steps with PBS, microscopy was carried out using an TCP. TZCP 1:1 showed the same peaks as TZCP 1:2. Moreover, additional
AX-10 fluorescence microscope (Axio Vision Imager M.1, Carl Zeiss peaks for monoclinic zirconia were also detected. On TZCP 2:1 charac-
AG, Jena, Germany). teristic peaks for α-TCP were observed, as well as monoclinic and
The metabolic activity of cells was determined with a colorimetric cubic zirconia. Thus, spectra of all sintered specimens showed phase de-
water-soluble tetrazolium salt (WST-1, Roche Diagnostics GmbH, composition of HA to TCP due to the high sintering temperature.
340 K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346

Fig. 1. SEM micrographs depicting the surface morphology of various coatings on TZ obtained by WPS after heat treatment at 1500 °C. A: pure CP coating, B: TZCP 1:2, C: TZCP 1:1, D: TZCP
2:1. B–D: light gray areas represent TZ and dark gray areas represent CP.

3.2. Coating stability and adhesion value was observed for pure CP and the highest value for the mixed
coating with the highest amount of zirconia (TZCP 2:1). TZCP 1:1 and
The biaxial flexural strength of the coated samples was determined TZCP 1:2 exhibited intermediate strength values indicating a linear de-
via B3B as shown in Fig. 3A. The mean strength values of CP, TZCP 1:2, pendence between biaxial flexural strength and amount of CP in the
TZCP 1:1 and TZCP 2:1 were 790 ± 55 MPa, 850 ± 63 MPa, 939 ± composition. All samples showed statistically significant differences.
64 MPa and 1023 ± 80 MPa, respectively (Fig. 3B). The lowest strength Investigations of fracture surfaces (Fig. 3C) support the B3B results.
TZCP 2:1 showed a smooth and continuous morphology along the
interbonding. Whereas CP, TZCP 1:2 and TZCP 1:1 exhibited a distinct
interface layer between substrate and coating, whose thickness in-
creased with enhanced CP content in the coating mixture (Fig. 3C
arrows).
Fig. 4 shows the sample surfaces before (A–D) and after scratch test
with a pencil (E–H) and a bovine femur (I–L). The black dashed lines
correspond to the pencil or bone scratch path. The results of the pencil
scratch test showed a completely damaged CP coating (Fig. 4E) that
was largely removed due to a poor coating adhesion. TZCP 1:2
(Fig. 4F) was only partially removed by the pencil indicating an inter-
mediate coating–substrate adhesion. In TZCP 1:1 (Fig. 4G) only minor
coating failures were observed that resulted solely in a pore enlarge-
ment. In contrast, TZCP 2:1 was completely resistant to the pencil sug-
gesting a firm attachment and a strong interbonding between coating
and substrate.
The bovine femur scratch test showed different results although the
pencil and bone exhibited the same Vickers hardness of about ~90 HV.
The pure CP coating (Fig. 4I) was again completely removed from the
substrate. Whereas all mixed coatings (Fig. 4J–L) demonstrated a good
mechanical integrity, no damage of the coating was observed after
bone burn out.
Zirconia screw implants (Fig. 5A) were tested with the same coating
compositions to evaluate the coating stability during implant place-
ment. Homogenously and fully coated implants were achieved by
WPS (Fig. 5B top row) and the coating morphology was similar to that
obtained on TZ discs. The implants were carefully dissected after inser-
Fig. 2. GI-XRD analysis of pure CP and various mixed coatings on zirconia after heat treat- tion into bovine rip bone (Fig. 5B, middle row). TZCP 2:1 exhibited a
ment at 1500 °C. high coating stability; the thread tips as well as the thread sides showed
K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346 341

Fig. 3. Mechanical strength behavior of various coated samples determined via B3B. A: Experimental set-up. B: Box plot diagram of the biaxial flexural strength. The black rectangle displays
the mean value and the horizontal line the median. All tested samples were significantly different from each other (p b 0.05). C: Corresponding fracture surfaces; the black arrows point the
interface between coating and substrate.

an intact coating confirming a strong adhesion between substrate and implants after dissection support these results (Fig. 5C bottom row).
coating, as no damages were detected. The other coatings possessed The SEM images depict coating failures after insertion particularly at
varying degrees of defects, especially at the thread tips. In areas with the thread tips. These findings are in agreement with the results from
no direct bone contact, the coating was still intact. Cross-sections of the scratch test (Fig. 3).

Fig. 4. SEM micrographs of the coating morphology before scratch test (A–D), after pencil scratch (E–H) and after bone scratch (I–L). The black dashed lines represent the scratch path.
342 K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346

Fig. 5. Coating adhesion and stability analysis of zirconia implant-like screws. A: Implant overview of a non-coated (left) and coated implant (right). B: SEM micrographs of the implant
coating morphology before insertion into bone (top row) and after dissection (middle row) with the appropriate cross section of the thread tips after dissection of bone and embedding in
resin (bottom row).

3.3. In vitro bioactivity assessment over time for all studied coatings followed by a gradual increase on
the following days. TZCP 1:2 showed the highest calcium ion release
Fig. 6 shows the cumulative calcium ion release in Tris–HCl solution followed by CP and TZCP 1:2. The lowest dissolution was recognized
represented as a function of time for all tested coatings. During the first for TZCP 2:1, indicating that Ca2+ release was predominantly decreased
4 days of immersion the ionic concentration of Ca2+ increased rapidly with increasing zirconia content.
SBF was used to investigate the biological activity of all studied
coatings in vitro, reflected in their capability to form a mineral
layer onto the surface. Fig. 7A shows the top view of all tested sam-
ples after 21 days in SBF. SEM micrographs of CP and TZCP 1:2 dem-
onstrate a completely covered surface by a newly formed mineral
layer with the typical globular morphology of apatite precipitates.
Cracks were also found in the mineral layer and are possibly formed
during the drying process. The higher magnification of the newly
formed layer on the CP surface evidenced a detailed layer structure
(inset in Fig. 7A). The surface of TZCP 1:2 was almost completely cov-
ered and TZCP 2:1, in contrast, showed only low mineral formation
ability. EDX analysis (Fig. 7B) of the mineral layer grown on the coat-
ed samples depicted the typical peaks of Ca, phosphor (P), magne-
sium (Mg), sodium (Na) and oxygen (O), with a Ca/P ratio of 2.3,
close to the theoretical value for apatite (2.15) [25]. However, the
carbon (C) peak is related to sputter coating with carbon. Fig. 7C
shows the Ca2 + concentration of the SBF supernatant on different
sampling days of all tested coatings. A significant decrease in Ca2 +
concentration over time was observed for all coatings in comparison
Fig. 6. Dissolution behavior of pure CP and various TZCP coatings sintered at 1500 °C after to TZCP 2:1. The reactivity of precipitate formation increased with
immersion at pH 7.4 for a period of 21 days as a function of immersion time. increasing CP content in the coating.
K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346 343

Fig. 7. Behavior of CP, TZCP 1:2, TZCP 1:1 and TZCP 1:2 coatings after immersion in simulated body fluid for 21 days at 37 °C. A: SEM micrographs of the newly formed layer on the tested
coatings with the typical apatite structure (inset), B: EDX elemental analysis of newly formed layer on the CP surface, C: Time dependent calcium ion concentration in the supernatant of
simulated body fluid solution at various sampling days.

3.4. In vitro biological characterization To evaluate the cell differentiation on the tested samples intracellu-
lar ALP activity of HOB cells was characterized for 4, 7 and 9 days
Immunofluorescence staining (Fig. 8) performed with human osteo- (Fig. 9D). In general, the ALP activity increased with time on all tested
blasts revealed cell morphology, spreading and growth after 1, 4 and coatings. ALP activities of cells cultured on CP and TZCP 2:1 were signif-
7 days of culture on all studied coatings. Cells cultured on the reference icantly pronounced in comparison to TZCP 1:2 and TZCP 1:1 coated
material Thermanox® were flattened and widespread already 1 day samples on all sampling days. As expected, the highest ALP activity
after seeding with many filopodia and well-organized actin fibers. In was detected on Thermanox®.
the following the cells grew further, until after 7 days of culture the sur-
face was completely covered by a dense and confluent cellular layer. On 4. Discussion
the coated samples the cells spread and grew to a different extent. After
1 day of culture the cells displayed a round or elongated shape. But on Pure HA or CP coatings on metallic dental implants were widely in-
TZCP 1:1 and TZCP 2:1 slightly more cells were found with a more flat- vestigated for an improved tissue integration because of their good
tened and spread morphology compared to CP and TZCP 1:2. At day 3 osteoconductivity, but these coatings are also known to have a poor sta-
the cells were spread well on all coatings except on TZCP 1:2 where bility. One of the major challenges in fabrication of coated implants is
most of the cells remained rounded and small. After 7 days of culture the maintenance of a suitable adhesion and stability during insertion ac-
proliferation was more pronounced on TZCP 2:1 and TZCP 1:1, indicat- companied with a favorable bioactivity. The present study demon-
ing better cell spreading and growth with enhanced content of TZ. strates the fabrication of mixed coatings with various ratios of CP and
WST-1 assay was used to determine HOB metabolic activity on all TZ. The coatings were applied with the WPS technique, which suitability
samples after days 1, 4, 7 and 9 as shown in Fig. 9A. The metabolic activ- was already shown in previous works [18,26]. During heat treatment
ity increased continuously during 9 days. After 1 day of culture, all test- the starting materials HA and TZ decomposed and transformed to
ed samples showed a nearly similar metabolic activity. At day 4 cells had other phases in all studied coatings due to the high sintering tempera-
a significant greater metabolic activity on TZCP 1:2 and TZCP 2:1 com- ture. The resulting phases were dependent on the initial TZ and HA con-
pared to CP and TZCP 1:2. However, the highest metabolic activity was tent, which either leads to a co-existence of β-TCP and α-TCP or to one
detected on Thermanox®. The same issue becomes more evident after of both phases. Furthermore, a transformation of tetragonal (t) to cubic
7 and 9 days of incubation. (c) zirconia caused by CaO diffusion [27,28] and the formation of mono-
Cell proliferation analyzed via the quantity of DNA (Fig. 9B) and pro- clinic zirconia were also detected and could compromise the implant
tein content (Fig. 9C) confirmed these observations. Within the coated strength. Therefore, understanding the mechanical as well as the biolog-
samples, the cell proliferation was retarded on TZCP 1:2, whereas the ical relevance in combination with the coating composition is of crucial
significant highest amount of cells was found on TZCP 2:1. In compari- importance.
son with the coated samples, Thermanox® exhibited a significantly Mechanical stability and coating adhesion were tested in this study
higher proliferation. by B3B, scratch test and dissection experiments of inserted zirconia
344 K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346

Fig. 8. Fluorescence micrograph of HOBs seeded on different coated samples over a culture period of 7 days compared to Thermanox®. The cells were fixed and stained with AF488-
phalloidin (green) for actin filaments and with DAPI for nuclei (blue).

implants. The results of the biaxial flexural strength test showed a major The adhesion depends on the coating–substrate bond strength and
impact of the coating composition on the mechanical strength. While plays an important role for the use of coated load-bearing implants.
for CP the flexural strength was strongly reduced, TZCP 2:1 nearly had Other studies have demonstrated the application of interfacial layers
no effect and is comparable with uncoated TZ, which has strength of to overcome the mismatch between coating and substrate due to differ-
around 1100 MPa [18]. The mechanical stability showed a linear depen- ent thermal expansion coefficients [29,30]. In this study a firm adhesion
dence between strength and coating composition, which significantly was reached by TZ addition to the coating and a coating–substrate co-
correlated with TZ:CP ratio. This behavior can be explained with in- sintering process in order to provide a sufficient bonding between coat-
creasing thickness of the interface layer between coating and substrate ing and substrate [28,31]. The adhesion was studied via qualitative
that increased with enhanced content of CP. During sintering released scratch test, performed with a pencil and bovine femur at constant
CaO from HA decomposition can also diffuse into zirconia substrates, load and speed. However, coated zirconia screws were inserted into
probably resulting in a transformation of tetragonal to cubic zirconia. fresh bovine rip bone, which is similar to jaw bone and characterized
With decreasing amount of TZ in the coating CaO is predominantly in- with respect to their coating integrity. The pure CP coating showed
corporated into the zirconia substrate and leads very likely to lower the weakest resistance in both experiments indicating a poor interfacial
strength values, because cubic zirconia has poor mechanical properties bonding between coating and substrate. The mixed coatings exhibited
compared to tetragonal zirconia [5]. different degrees of damage, which were more pronounced with
K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346 345

Fig. 9. A) Quantification of metabolic activity, (B) cell proliferation, (C) protein content and (D) ALP activity of HOB cultured on different coated samples and on Thermanox® over a culture
period of 9 days. Different symbols above the columns mark the significance level (p b 0.05) tested by ANOVA.

increased CP content. In other words, the higher the TZ content the Bioactive properties of material surfaces play a crucial role in the
stronger the interbonding between coating and substrate. Slightly dif- evaluation of cell-biomaterial interactions in order to assess the efficacy
ferent results obtained from the pencil and bone scratch test can prob- of their application in human medicine. Therefore, cellular response of
ably be explained by different stiffness and elasticity characteristics of HOBs, such as cell growth, proliferation and differentiation was investi-
the scratching materials. These results show that the scratch test can gated in order to assess the influence of surface chemistry and physical
give a first indication of the coating stability. But insertion and dissec- features. Fluorescence microscopic images revealed that cells on coat-
tion experiments, which are similar to a clinical situation, provide ings with increasing TZ contents showed faster attachment, a more
more reliable data. spread-out morphology, taking a polygonal shape with several cyto-
Dissolution experiments showed that the Ca2 + release was very plasmic extensions and an apparently higher cell density. Higher CP
similar for all samples at the beginning of the experiment and differ contents, in contrast, seemed to hamper cell flattening and spreading.
only slightly in the further course of time. TZCP 1:2 exhibited the Metabolic activity, cell number and protein content were significantly
highest Ca ion release and TZCP 2:1 the lowest. However, the trend of greater on TZCP 1:1 and TZCP 2:1 at days 4–9 compared to the other
the cumulative release curves is the same for all tested specimens. The coatings, indicating a good cytocompatibility. ALP activity followed
dissolution behavior is directly influenced not only by the phase compo- almost the same trend with the exception of an elevated value on CP.
sition on the one hand, but also by the microstructure on the other hand. These results show that the cells respond differently to the investi-
With increased TZ content in the coating less CP is directly exposed at gated coatings, indicating an influence of the coating properties.
the surface as can be seen in Fig. 1, thus less calcium and phosphate The highest cell proliferation, viability and differentiation were
can dissolve [32]. found on Thermanox®. The cells were flattened, well spread and
SBF bioactivity results indicated significant mineral layer formation reached confluence at day 9, which is due to the sample's physical
abilities. After 21 days of incubation in SBF solution, all coatings showed properties and consistent with the work of others [34]. These
a newly formed layer with typical cauliflower morphology except for in vitro investigations showed that all studied coatings were non-
TZCP 2:1. The corresponding EDX analyses revealed calcium and phos- toxic. Nevertheless, due to similar coating surface topography and
phorus as main components, as well as magnesium and sodium in roughness, the varying extent of cell behavior can be mainly described
small amounts, which might have similarities to bone apatite. Changes by different coating dissolution. Coatings with a higher amount of CP
of Ca2+ concentration in SBF revealed the tendency for mineralization dissolve more rapidly and thus lead to an alkaline pH of the cell culture
processes with increasing content of CP in the coating. The higher the medium which affected the initial cell viability [35,36].
mineral forming ability, the lower the Ca2+ concentration in SBF solu- Nevertheless, further work is clearly needed to characterize the
tion was. Thus, precipitation of bone-like minerals reflects the dissolu- in vivo bioactivity of the coating, as well as the long-term stability. Con-
tion behavior of the studied coatings. The same observation was made sequently, it is imperative to know the coating stability after insertion
by Weng et al. who showed that no precipitation occurred on surfaces and the dissolution rate in a living organism under physiological condi-
with low dissolution [33]. tions. Moreover, the suspension composition can be altered, by
346 K. Pardun et al. / Materials Science and Engineering C 48 (2015) 337–346

introducing further additives, in order to get closer to the composition [13] L.L. Hench, Bioceramics: from concept to clinic, J. Am. Ceram. Soc. 74 (1991)
1487–1510.
of natural bone. [14] X. Liu, P. Chu, C. Ding, Surface modification of titanium, titanium alloys, and related
materials for biomedical applications, Mater. Sci. Eng. R. Rep. 47 (2004) 49–121.
[15] E. Mohseni, E. Zalnezhad, A.R. Bushroa, Comparative investigation on the adhesion
5. Conclusion
of hydroxyapatite coating on Ti–6Al–4V implant: a review paper, Int. J. Adhes.
Adhes. 48 (2014) 238–257.
The present study examined the influence of different TZ and CP ra- [16] S. Yugeswaran, C.P. Yoganand, a. Kobayashi, K.M. Paraskevopoulos, B. Subramanian,
tios on the mechanical and biological properties of coatings on zirconia Mechanical properties, electrochemical corrosion and in-vitro bioactivity of yttria
stabilized zirconia reinforced hydroxyapatite coatings prepared by gas tunnel type
substrates. Coating adhesion and mechanical stability were consider- plasma spraying, J. Mech. Behav. Biomed. Mater. 9 (2012) 22–33.
ably increased with increasing addition of TZ. SBF test results, however, [17] T.J. Matsumoto, S.-H. An, T. Ishimoto, T. Nakano, T. Matsumoto, S. Imazato, Zirconia–
revealed a higher bioactivity potential of coatings with enhanced CP hydroxyapatite composite material with micro porous structure, Dent. Mater. 27
(2011) e205–e212, http://dx.doi.org/10.1016/j.dental.2011.07.009.
content due to different dissolution behaviors. Despite similar surface [18] K. Pardun, L. Treccani, E. Volkmann, G. Li Destri, G. Marletta, P. Streckbein, et al.,
morphology and average roughness in vitro investigations showed Characterization of wet powder-sprayed zirconia/calcium phosphate coating for
that the cells respond differently with regard to proliferation and differ- dental implants, Clin. Implant. Dent. Relat. Res. (2013), http://dx.doi.org/10.1111/
cid.12071.
entiation, most likely affected by dissolution of the studied coatings. [19] I. Lhermitte-Sebire, R. Colmet, R. Naslain, J. Desmaison, G. Gladel, The adhesion be-
Considering the overall good mechanical strength, the excellent interfa- tween physically vapour-deposited or chemically vapour-deposited alumina and
cial bonding and the bioactivity potential, TZCP 1:1 and TZCP 2:1 pro- TiC-coated cemented carbides as characterized by Auger electron spectroscopy
and scratch testing, Thin Solid Films 138 (1986) 221–233.
vide a highly interesting coating for dental implants. [20] L. Montanaro, C.R. Arciola, D. Campoccia, M. Cervellati, In vitro effects on MG63
osteoblast-like cells following contact with two roughness-differing
Acknowledgments fluorohydroxyapatite-coated titanium alloys, Biomaterials 23 (2002) 3651–3659.
[21] M. Stefanic, K. Krnel, T. Kosmac, Novel method for the synthesis of a β-tricalcium
phosphate coating on a zirconia implant, J. Eur. Ceram. Soc. 33 (2013) 3455–3465.
Authors would like to acknowledge financial support from the [22] D. Barnes, S. Johnson, R. Snell, S. Best, Using scratch testing to measure the adhesion
Deutsche Forschungsgemeinschaft, DFG (Grant No. RE 2735/7-1). We strength of calcium phosphate coatings applied to poly(carbonate urethane) sub-
strates, J. Mech. Behav. Biomed. Mater. 6 (2012) 128–138.
thank Petra Witte of Historical Geology–Palaeontology, University of [23] A. Börger, P. Supancic, R. Danzer, The ball on three balls test for strength testing of
Bremen for the performance of SEM and EDX analysis. We are also brittle discs: stress distribution in the disc, J. Eur. Ceram. Soc. 22 (2002) 1425–1436.
grateful to Martin Ellerhorst from BEGO Implant Systems for providing [24] M.G. Holthaus, J. Stolle, L. Treccani, K. Rezwan, Orientation of human osteoblasts on
hydroxyapatite-based microchannels, Acta Biomater. 8 (2012) 394–403.
implant geometries and Andreas Reindl from Fraunhofer IFAM Bremen [25] E.M. White, L.A. Hannus, Chemical weathering of bone in archaeological soils, Am.
for preparation of zirconia implants, as well as Michael Teck for his help Antiq. 48 (1983) 316.
with experimental investigations. [26] A. Ruder, H.P. Buchkremer, H. Jansen, W. Malléner, D. Stöver, Wet powder spraying
— a process for the production of coatings, Surf. Coat. Technol. 53 (1992) 71–74.
[27] K. Ioku, M. Yoshimura, S. Somiya, Microstructure and mechanical properties of hy-
References droxyapatite ceramics with zirconia dispersion prepared by post-sintering, Bioma-
terials 11 (1990) 57–61.
[1] B. Reinhardt, T. Beikler, Dental implants, in: J.Z. Shen, T. Kosmač (Eds.), Adv. Ceram. [28] L. Fu, K.A. Khor, J.P. Lim, Effects of yttria-stabilized zirconia on plasma-sprayed
Dent, Elsevier Inc., Amsterdam, 2014, pp. 51–76. hydroxyapatite/yttria-stabilized zirconia composite coatings, J. Am. Ceram. Soc. 85
[2] M. Niinomi, Mechanical properties of biomedical titanium alloys, Mater. Sci. Eng. A (2002) 800–806.
243 (1998) 231–236. [29] J.-Z. Yang, R. Sultana, X.-Z. Hu, Z.-H. Huang, Porous hydroxyapatite coating on strong
[3] H. Tschernitschek, L. Borchers, W. Geurtsen, Nonalloyed titanium as a bioinert metal ceramic substrate fabricated by low density slip coating-deposition and coating–
— a review, Quintessence Int. 36 (2005) 523–530. substrate co-sintering, J. Eur. Ceram. Soc. 31 (2011) 2065–2071.
[4] C. Wu, Y. Ramaswamy, D. Gale, W. Yang, K. Xiao, L. Zhang, et al., Novel sphene coat- [30] H.-W. Kim, H.-E. Kim, V. Salih, J.C. Knowles, Dissolution control and cellular re-
ings on Ti–6Al–4V for orthopedic implants using sol–gel method, Acta Biomater. 4 sponses of calcium phosphate coatings on zirconia porous scaffold, J. Biomed.
(2008) 569–576. Mater. Res. A. 68 (2004) 522–530.
[5] C. Piconi, G. Maccauro, Zirconia as a ceramic biomaterial, Biomaterials 20 (1999) [31] J. Zhang, M. Iwasa, N. Kotobuki, T. Tanaka, M. Hirose, H. Ohgushi, et al., Fabrication of
1–25. hydroxyapatite–zirconia composites for orthopedic applications, J. Am. Ceram. Soc.
[6] J. Chevalier, What future for zirconia as a biomaterial? Biomaterials 27 (2006) 89 (2006) 3348–3355.
535–543. [32] M. Stefanic, R. Milacic, G. Drazic, M. Skarabot, B. Budič, K. Krnel, et al., Synthesis of
[7] P.F. Manicone, P. Rossi Iommetti, L. Raffaelli, An overview of zirconia ceramics: basic bioactive β-TCP coatings with tailored physico-chemical properties on zirconia
properties and clinical applications, J. Dent. 35 (2007) 819–826. bioceramics, J. Mater. Sci. Mater. Med. 25 (2014) 2333–2345.
[8] G. Balasundaram, T.J. Webster, Increased osteoblast adhesion on nanograined Ti [33] J. Weng, Q. Liu, J.G. Wolke, X. Zhang, K. de Groot, Formation and characteristics of
modified with KRSR, J. Biomed. Mater. Res. A 80A (2007) 602–611. the apatite layer on plasma-sprayed hydroxyapatite coatings in simulated body
[9] L. Le Guéhennec, A. Soueidan, P. Layrolle, Y. Amouriq, Surface treatments of titanium fluid, Biomaterials 18 (1997) 1027–1035.
dental implants for rapid osseointegration, Dent. Mater. 23 (2007) 844–854. [34] M.G. Holthaus, L. Treccani, K. Rezwan, Osteoblast viability on hydroxyapatite with
[10] V. Borsari, G. Giavaresi, M. Fini, P. Torricelli, A. Salito, R. Chiesa, et al., Physical char- well-adjusted submicron and micron surface roughness as monitored by the prolif-
acterization of different-roughness titanium surfaces, with and without hydroxyap- eration reagent WST-1, J. Biomater. Appl. 27 (2013) 791–800.
atite coating, and their effect on human osteoblast-like cells, J. Biomed. Mater. Res. B [35] Q. Tang, R. Brooks, N. Rushton, S. Best, Production and characterization of HA and
Appl. Biomater. 75 (2005) 359–368. SiHA coatings, J. Mater. Sci. Mater. Med. 21 (2010) 173–181.
[11] S.V. Dorozhkin, Calcium orthophosphates in dentistry, J. Mater. Sci. Mater. Med. 24 [36] I. Izquierdo-Barba, F. Conde, N. Olmo, M.a. Lizarbe, M.a. García, M. Vallet-Regí, Vitre-
(2013) 1335–1363. ous SiO2–CaO coatings on Ti6Al4V alloys: reactivity in simulated body fluid versus
[12] M. Jarcho, Calcium phosphate ceramics as hard tissue prosthetics, Clin. Orthop. osteoblast cell culture, Acta Biomater. 2 (2006) 445–455.
Relat. Res. (1981) 259–278.

You might also like