• Aucun résultat trouvé

Effects of LP-MOCVD prepared TiO2 thin films on the in vitro behavior of gingival fibroblasts

N/A
N/A
Protected

Academic year: 2021

Partager "Effects of LP-MOCVD prepared TiO2 thin films on the in vitro behavior of gingival fibroblasts"

Copied!
9
0
0

Texte intégral

(1)

O

pen

A

rchive

T

OULOUSE

A

rchive

O

uverte (

OATAO

)

OATAO is an open access repository that collects the work of Toulouse researchers and

makes it freely available over the web where possible.

This is an author-deposited version published in :

http://oatao.univ-toulouse.fr/

Eprints ID : 4665

To link to this article : DOI :10.1016/j.matchemphys.2010.10.028

URL :

http://dx.doi.org/10.1016/j.matchemphys.2010.10.028

To cite this version : Cimpean, A. and Popescu, Simona and

Ciofrangean, C.M. and Gleizes, Alain N. ( 2011) Effects of

LP-MOCVD prepared TiO2 thin films on the in vitro behavior of

gingival fibroblasts. Materials Chemistry and Physics, vol. 125

(n° 3). pp.485-492. ISSN 0254-0584

Any correspondance concerning this service should be sent to the repository

administrator: staff-oatao@inp-toulouse.fr

.

(2)

Effects of LP-MOCVD prepared TiO

2

thin films on the in vitro

behavior of gingival fibroblasts

Anisoara Cimpean

a

, Simona Popescu

b,c

, Cristina M. Ciofrangeanu

a

, Alain N. Gleizes

c,∗

aDepartment of Biochemistry and Molecular Biology, University of Bucharest, Bucharest, Romania

bFaculty of Applied Chemistry and Materials Science, University Politehnica of Bucharest, Polizu, Bucharest, Romania

cCentre Interuniversitaire de Recherche et d’Ingénierie des Matériaux (CIRIMAT), CNRS-INPT/ENSIACET, University of Toulouse, Toulouse, France

Keywords: Biomaterials TiO2

Chemical vapor deposition Fibroblast cell culture

a b s t r a c t

We report on the in vitro response of human gingival fibroblasts (HGF-1 cell line) to various thin films of titanium dioxide (TiO2) deposited on titanium (Ti) substrates by low pressure metal-organic chemical

vapor deposition (LP-MOCVD). The aim was to study the influence of film structural parameters on the cell behavior comparatively with a native-oxide covered titanium specimen, this objective being topical and interesting for materials applications in implantology. HGF-1 cells were cultured on three LP-MOCVD pre-pared thin films of TiO2differentiated by their thickness, roughness, transversal morphology, allotropic

composition and wettability, and on a native-oxide covered Ti substrate. Besides traditional tests of cell viability and morphology, the biocompatibility of these materials was evaluated by fibronectin immunos-taining, assessment of cell proliferation status and the zymographic evaluation of gelatinolytic activities specific to matrix metalloproteinases secreted by cells grown in contact with studied specimens. The ana-lyzed surfaces proved to influence fibronectin fibril assembly, cell proliferation and capacity to degrade extracellular matrix without considerably affecting cell viability and morphology. The MOCVD of TiO2

proved effective in positively modifying titanium surface for medical applications. Surface properties playing a crucial role for cell behavior were the wettability and, secondarily, the roughness, HGF-1 cells preferring a moderately rough and wettable TiO2coating.

1. Introduction

Nowadays, biomaterials are used in a different way compared to just a decade ago. Despite this fact, implantable medical devices are still important, and medical technologies and tissue engineer-ing still encompass bioinert metals, activated or/and bioactivated, taking into account that expectations of their performance also have been changed[1]. In this context, Ti and Ti alloys are the most commonly used metal materials to manufacture orthope-dic and dental implants owing to its excellent physical, chemical, mechanical properties, and biocompatibility with receiving liv-ing tissues after implantation [2]. An essential surface feature of Ti is the capacity to spontaneously form a stable amorphous TiO2 film under exposure to the atmosphere and/or

physiolog-ical fluids. This passive, 4–6 nm thick film[3] protects Ti from corrosion, delays the release of Ti ions[4,5], and promotes a favor-able osseointegration. In dentistry, the passive TiO2 film formed

on titanium seems to be more stable and protective than that

∗ Corresponding author at: ENSIACET, 4 allée Emile Monso, BP-44362, 31432, Toulouse Cedex 4, France. Tel.: +33 5 34 32 34 39; fax: +33 5 34 32 34 98.

E-mail address:alain.gleizes@ensiacet.fr(A.N. Gleizes).

formed on the Ti alloys, usually used in other medical applica-tions.

The evolution of biocompatibility concept[6]is based on the idea that the response to specific individual materials could vary from one application to another, biocompatibility being dependent on the material characteristics but only in a specific application. In such a way, dental implants are unique in the body when com-pared to other medical implant types, in that they create two interfaces: the implant bone interface (at the radicular portion of the implant) and the gingival tissue implant interface. The main-tenance of this last interface seems to be as important as bone maintenance adjacent to the implant. The area of gingival tissue at the point where the natural tooth emerges from the bone has to regenerate around the neck of an implant in order to protect the underlying connective tissue and to prevent the implant from moving and loosening eventually. Thus, the response of gingival epithelial and connective tissue cells to implant surface is very important.

The long-term success of implants most importantly requires stable interfacial interactions between biomaterial and the surrounding tissue. Therefore, different surface modification approaches have been attempted to improve the bioactive bone-binding capacity of titanium metal[6–13]. They have been applied

(3)

especially to form a bioactive TiO2 layer on the metal

sur-face.

A technique that has scarcely been applied to the deposition of titanium dioxide for clinical purposes is metal-organic chemi-cal vapor deposition (MOCVD)[14–16]. This technique is widely used to prepare TiO2thin films at relatively low temperature. It is

well adapted to the coating of complex-shaped surfaces, especially at low pressure. The microstructural parameters (crystallinity, growth mode, grain size, texture, roughness, allotropic composi-tion, surface energy. . .) of the deposit prepared by MOCVD sensibly depends on the numerous experimental parameters of the tech-nique, mainly: surface state of the substrate, growth temperature, total pressure, partial pressure and molar fraction of the precursor, time of deposition. Therefore varied surface microstructures can be obtained just by playing with the parameters. The MOCVD tech-nique thus offers the possibility to study the influence of various surface structure parameters on the biocompatibility of cells.

To our knowledge, the few studies devoted to the elucidation of in vitro interactions of the cells with TiO2in different physical

states deal with TiO2particles blasted on Ti substrates[17], oxide

layers prepared by MOCVD[11,14–16]or anodic oxidation[18,19], cluster-assembled nanostructured TiO2 films [20], and

sol–gel-derived titanium oxide[21,22].

In this context, the present study was undertaken to evaluate the biocompatibility of various TiO2thin films deposited on titanium

substrates using MOCVD technique. Biocompatibility was evalu-ated in culture of human gingival fibroblasts, by addressing cell viability, cytomorphology, adhesive cell-biomaterial interactions mediated through fibronectin (FN) matrix, proliferation capacity and potential for degradation of the extracellular matrix (ECM), as indicators of the biological performance of studied materials. 2. Experimental

2.1. Preparation of materials

The substrates were pure Ti discs of 1 mm thickness and 10 mm diameter (99.6% purity, Good Fellow Ltd.). Deposition surfaces were mechanically polished with 180-grit SiC-paper. Before TiO2deposition, the discs were ultrasonically cleaned

in acetone, then in boiling alcohol, and rinsed with distilled water. Control discs were cleaned in the same way and used as covered with their native oxide. TiO2

thin films were grown using LP-MOCVD with titanium tetra-isopropoxide (TTIP) as a precursor, and nitrogen as both carrier and dilution gas. The procedure and the experimental set-up were described in a previous paper[10].

Deposition parameters were: TTIP temperature, 25◦C; N2flow in the precursor

container, 20 sccm; dilution N2flow, 575 sccm; total pressure, 20 Torr; TTIP molar

fraction in the gas phase, 76× 10−6; deposition temperature, 400C; deposition time

ranging from 90 to 360 min to vary deposit thickness.

2.2. Characterization of TiO2films

The three LP-MOCVD TiO2-coated Ti samples prepared for this study are

char-acterized by the following parameters: (1) film thickness, deduced from the weight of deposited TiO2; (2) film transversal morphology, examined with a LEO-435

scan-ning electron microscope (SEM) operating at 15 kV; (3) roughness (Ra), measured on

0.5 mm2of samples surface area with an optical interferometer MetroProTM, Zygo

New View 100 equipment; (4) water contact angles measured with a DIGIDROP Contact Angle Meter (GBX Scientific Instruments).

Deposited phases were identified from X-ray diffraction patterns recorded with a vertical, theta-theta geometry diffractometer (Seifert XRD 3000 TT) equipped with a graphite monochromator and a Cu anticathode ( CuK␣1= 1.5406 ˚A). Both grazing incidence X-ray diffraction (GIXRD) patterns (grazing angle of 2◦) and– patterns were recorded.

2.3. Cells and cell culture conditions

HGF-1 cells (American Type Culture Collection) were seeded on TiO2-coated and

native-oxide covered Ti discs, placed into Permanox Lab-Tek Chamber Slides (Nunc), at an initial density of 1.2× 104cells cm−2. They were grown in a Dulbecco’s Modified

Eagle Medium (DMEM) containing 10% (v/v) fetal bovine serum (FBS) at 37◦C in a

standard incubator with 5% CO2in air. Prior to cell culture, all these materials were

sterilized at 180◦C for 30 min. All experiments were repeated 3 times.

2.4. Assay of cell viability

The eventual toxic effects of the analyzed specimens on cultured gingival fibrob-lasts were evaluated by detecting lactate dehydrogenase (LDH) activity released into the media as a marker of cell death and plasma membrane lysis. This study was per-formed in triplicate samples from the cell culture media maintained in contact with TiO2-coated and native-oxide covered Ti discs for 72 h, by using a cytotoxicity

detec-tion kit (TOX-7, Sigma–Aldrich Co.) according to the manufacturer’s protocol. High OD490 nmvalues are indicative of a reduction in the cell viability.

2.5. Immunocytochemical staining of actin and fibronectin

At the end of the incubation period (72 h and, respectively, 120 h for studies of actin and fibronectin expression), cells were fixed with 4% paraformaldehyde in PBS for 20 min and observed by phase contrast microscopy. Staining of actin was performed with fluorescein isothiocyanate (FITC)–phalloidin (20␮g mL−1;

Sigma–Aldrich Co.) at 37◦C for 1 h. Before this, they were permeabilized for 15 min

with 0.1% Triton X-100/2% BSA and washed in PBS. For FN visualization, fixed and permeabilized specimens were incubated with anti-FN (mouse) primary antibody (EP5 clone from Santa Cruz Biotechnology, dilution 1:50) for 2 h, and then incubated with anti-mouse secondary antibody labeled with FITC (Santa Cruz Biotechnology, dilution 1:200) for 1 h. In this case, the nuclei were counterstained with DAPI to reveal cell density.

All the Ti specimens were viewed with an inverted fluorescent microscope Olympus IX71. Excitation wavelengths were 495 nm for FITC and 358 nm for DAPI. The images were captured by means of Cell F image acquiring system.

2.6. BrdU immunocytochemistry

After 6 h treatment of HGF-1 cells (at 24 h post-seeding) with 5-bromo-2

-deoxyuridine (BrdU) at a final concentration of 10␮M, TiO2-coated and native-oxide

covered Ti discs were maintained in 4% paraformaldehyde for 20 min. To enable antibody binding to the incorporated BrdU into the DNA of dividing cells, cells were permeabilized with 1% Triton X-100 and the DNA was denatured by sequential treat-ment with 1N HCl (20 min on ice) and 2 N HCl (10 min at room temperature and then 40 min at 37◦C). After washing in borate buffer (0.1 M, pH 8.5) and blocking with 1% Triton X-100/5% BSA, samples were incubated for 2 h with a monoclonal mouse anti-BrdU antibody (Sigma–Aldrich Co.) diluted 1:100. Antibody binding was detected by incubating discs for 1 h at room temperature with FITC-conjugated goat anti-mouse IgG (Santa Cruz Biotechnology) used at 1:100 dilution. After washing with PBS, samples were examined under an inverted fluorescent microscope Olympus IX71.

In this study and in the immunocytochemical assays, previously presented, neg-ative controls were realized by substitution of the primary antibodies with PBS.

At least 10 fields from three independent experiments were examined. 2.7. Western blot assay

Total cell lysates containing 30␮g protein, as determined by Bradford method [23], were separated by SDS-PAGE on 10% polyacrylamide gels. After electrophore-sis, proteins from the gel were transferred to nitrocellulose membranes (Invitrogen 0.45␮m pore nitrocellulose/filter paper sandwich; XCell IITMBlot Module).

Mem-branes were blocked with 5% nonfat milk, and probed with proliferating cell nuclear antigen (PCNA) mouse monoclonal antibody (1:200) for at least 2 h. In order to detect the antibodies which have bound, the membranes were incubated with alkaline phosphatase-conjugated goat anti-mouse IgG (1:7500) for 1 h, and developed with NBT/BCIP solution. The blots were reprobed with anti-␣-actin antibody to confirm equal protein loading. Both primary and secondary antibodies were purchased from Santa Cruz Biotechnology.

2.8. Gelatin zymography

Gelatin zymography was performed by using a modification of the procedure by Cimpean et al.[24]. Briefly, aliquots (20␮g protein) of conditioned culture medium, maintained in contact for 72 h with the sub-confluent cultures, were subjected to SDS-PAGE in 7.5% (w/v) polyacrylamide gels impregnated with 1 mg mL−1gelatin

under non-reducing conditions.

After electrophoresis, gels were washed two times for 30 min each in a buffer containing 50 mM Tris–HCl (pH 7.6), 50 mM NaCl, 10 mM CaCl2, 0.05% Brij 35 and

2.5% Triton X-100 and then incubated for 18 h at 37◦C with the same buffer excluding Triton X-100. After incubation, the gels were stained with 0.1% Coomassie Brilliant Blue-R 250 and then destained in acetic acid: methanol: water (1:2:7). Matrix met-alloproteinase (MMP) specific gelatinolytic activities were detected as clear bands against a background of blue stained, intact gelatin impregnated acrylamide gel. 2.9. Statistical analysis

Statistical analysis of surface roughness and the LDH assay was performed with GraphPrism software using one-way ANOVA with Bonferroni’s multiple comparison tests.

(4)

Table 1

A summary of TiO2films properties.

Specimen code T-1200 T-9 T-12 T-14 Deposition time (min) No 90 250 360 Thickness (nm) A few nm thin native oxide layer 200 600 1000 Transversal morphology Compact Compact Compact undercoat + thin columnar overlay Columnar Roughness Ra(␮m)a 0.35 (0.02) 0.71 (0.01) 0.75 (0.04) 0.92 (0.14)

Contact angle (◦) 60 55 9 8

Phasesb Amorphous A–R a–R A–R

aStandard deviations are given in parentheses. Mean values are statistically significantly different P < 0.01 (T9 versus T1200) and P < 0.001 (T12 versus T1200; T14 versus

T1200).

bA (a) = majority (minority) of anatase; R (r) = majority (minority) of rutile.

3. Results and discussion

3.1. Characterization of TiO2films

It is well known that host tissues specifically interact with the outermost layer of an implant[25]. Thus, when a Ti implant

is inserted into the human body, the surrounding tissues come into direct contact with the protective and bioactive oxide layer, mainly represented by TiO2. Considering that the oxide layer

which forms naturally on titanium is only a few nanometers thick and presents imperfections, we used the LP-MOCVD technique to deposit TiO2 layers on flat Ti substrates with the purpose of

Fig. 1. Surface micrographs of (a) native-oxide covered Ti surface, and MOCVD-grown TiO2films on Ti substrates for different periods of time: (b) 90 min; (c) 250 min, and

(5)

Fig. 2. GIXRD patterns of the TiO2films deposited on Ti substrate.

improving the interfacial properties between the gingival tissue and the implant. These samples were called T9, T12 and T14 and they are presented inTable 1 along with their preparative and structural characteristics. T1200 is a native-oxide covered Ti specimen (Fig. 1a) which was polished with grade 1200 SiC paper prior to use as a control in cell culture experiments.

SEM micrographs evidenced different film morphologies from small (T9:Fig. 1b) to large (T12:Fig. 1c, and T14:Fig. 1d) grained deposits, depending on the deposition time: the longer the time, the thicker the film and the larger the crystal size. Big crystal clus-ters grown here and there over the surface increase the surface roughness. Even on rough surfaces the films proved conformal, i.e. they fit in polishing grooves.

Fig. 1e and f shows the cross sectional morphology of samples T9 and T14. The 1␮m thick deposit on T14 shows a columnar structure, whereas the 0.2␮m thick deposit on T9 shows a compact struc-ture. This is consistent with a previous study showing that under the growth conditions listed in Section2.1, the TiO2deposit adopts

a compact structure as long as the thickness is less than 0.4␮m or so, and that a columnar structure forms upon increasing the thick-ness[26]. T12 is 0.6␮m thick. Its cross sectional morphology (not shown onFig. 1) is made of a compact undercoat covered with a thin columnar overlay. The measured water contact angles (Table 1) are consistent with the observed morphologies. The high hydrophilic character of samples T12 (contact angle = 9◦) and T14 (8◦) comes from the micro-porous surface generated by inter columnar spac-ing. Conversely, the compact structure of T9 results in a far less hydrophilic surface with a water contact angle of 55◦.

From GIXRD patterns, the films grown on T9, T12 and T14 were mixtures of crystallized anatase and rutile (Fig. 2). The respective amounts of anatase and rutile could not be measured quantitatively from the– patterns, either because intensities were too weak due to the thinness of deposit (T9) or because rutile proved2 0 0 textured (T12 and T14). This is whyTable 1shows but qualitative evaluations.

3.2. Cell viability and cell morphology

An important objective of this study was to determine if the thin films of TiO2produced by LP-MOCVD technique affected

cel-lular survival. To test this parameter, the LDH activity in the culture media was quantified. LDH is a soluble cytosolic enzyme, present in all cell types, and rapidly released into the cell culture medium upon damage of the plasma membrane. Therefore, it is the most widely used marker of cell membrane integrity and serves as a general means to assess cytotoxicity. HGF-1 cells in contact with TiO2-coated and native-oxide covered Ti surfaces proved to

dis-play low LDH release in culture medium suggesting that none of these surfaces exerted cytotoxic effects. Moreover, no significant differences between the samples were noticed (Table 2).

In culture, the morphology indicates the health status of the cells. As shown inFig. 3, at 72 h post-seeding, HGF-1 cells were fully spread, elongated and polygonal shaped, with cytoplasmic extensions and filopodia, indicating good cell attachment to the surface. Cell body of the gingival fibroblasts cultured on MOCVD-coated surfaces (Fig. 3b–d) became increasingly more voluminous with increased roughness. This could be due to a sparse adsorption of proteins from culture medium on the surface of oxide crystals. These protein aggregates represent interaction sites to cells that, in this situation, extend across large areas adopting a more elongated and large morphology. On the four surfaces, an intense labeling of actin cytoskeleton was observed in the entire cytoplasm and bright fluorescent filiform stress fibers were visible.

3.3. Adhesive cell-biomaterial interactions mediated through fibronectin matrix

Cell adhesion to a biomaterial surface is, in many cases, a spe-cial form of cell–ECM interaction, since the biomaterial surface is always covered by proteins that become adsorbed to this from either serum within cell culture media, or body fluid within the in vivo environment. Some of these proteins with which the cells interact, such as fibronectin and vitronectin, represent very known ligands for the integrin family of cell adhesion receptors [27]. Integrin-mediated cell adhesion is associated with signaling events and modulation of gene expression that ultimately determine the following cellular events: proliferation, differentiation and survival

[28]. Inadequate or inappropriate cell–ECM adhesions can lead to anoikis, a special type of adhesion driven apoptosis[29].

Fibronectin is an abundant component of the ECM consisting of repeating units of amino acids, which form domains that enable the molecule to interact with many cell types through both inte-grin and non-inteinte-grin receptors. It is encoded by a single gene, but alternative splicing of pre-mRNA allows formation of multi-ple isoforms that regulate adhesion-dependent survival signaling

[30]and play a critical role in cell proliferation and differentiation

[28]. FN binds primary to the␣5␤1 integrins, through the arginine-glycine-aspartic acid (RGD) cell-binding site forming a prototypic integrin–ligand pair[31]. This receptor–ligand pair mediates FN fibril formation and governs ECM assembly, which is essential to cell function. During assembly, FN is initially organized into fine cell-associated fibrils and, through continued accumulation of FN by synthetic and secretory cell activities, these fibrils are organized into a dense and stable extracellular fibrillar network. Therefore, fibronectin organization could be a useful tool to determine the biocompatibility of dental implants. After 5 days of culture, we found out a different distribution pattern of this protein over ana-lyzed surfaces (Fig. 4). Thus, removal and reorganization of FN

Table 2

Comparative evaluation of LDH activity expressed in culture media by HGF-1 cells in contact with TiO2coated (T9, T12, T14) and native-oxide covered

(T1200) specimens. Data show absorbance values (OD490) averaged from three measurements by specimen. Experimental errors are expressed as three

times the standard deviations. Mean values are not statistically significantly different P > 0.05 (T9, T12 and T14 versus T1200).

Specimen Ti-1200 T9 T12 T14 OD490 0.251 (0.003) 0.249 (0.001) 0.254 (0.001) 0.258 (0.001)

(6)

Fig. 3. Morphology of HGF-1 cells attached to: (a) native-oxide covered Ti surface, and MOCVD-TiO2 coated specimens: (b) T9, (c) T12, and (d) T14. Fluorescent actin staining

of cytoskeleton (FITC-conjugated phalloidin).

from the material surfaces into ECM like structures occurred on moderately wettable surfaces T1200 (Fig. 4a) and T9 (Fig. 4b). How-ever, fibrillar FN network was better expressed on T9 than on the non-CVD-coated T1200 sample. The latter shows less

fibril-lar structures and veil-like FN zones indicating a non-uniformous removal/reorganization of FN adsorbed to and secreted on the material surface. On the more hydrophilic surfaces (T12 and T14), the background was majoritary FN immunostained and very few

Fig. 4. Indirect immunolocalization of fibronectin adsorbed to and secreted by HGF 1 cells on: (a) native-oxide covered Ti surface, and MOCVD-TiO2coated specimens: (b)

(7)

Fig. 5. Immunofluorescent detection of proliferative cells after BrdU incorporation showing the cells that undergo DNA synthesis, as indicated by presence of green bright nuclei, on: (a) native-oxide covered Ti surface, and MOCVD-TiO2coated specimens: (b) T9, (c) T12, and (d) T14.

fibrils assembled (Fig. 4c and d). This suggests that the cells were essentially unable to reorganize FN into an ECM-like structure on the material surface.

Our results confirm the influence of surface wettability on the interactions of the cellular substrate with proteins and fur-ther on cell behavior. It is a general trend that mammalian cells interact better with hydrophilic (wettable) surfaces than with hydrophobic (non-wettable) surfaces[32–34]. In fact, the wetta-bility of the surface may affect cell attachment either directly, since the attachment phase as an initial process involves physic-ochemical interactions between cells and surfaces including ionic forces or indirectly through alterations in the adsorption pro-cess of conditioning molecules, such as proteins. On the contrary, other authors reported that cells adhere better on moderately hydrophobic surfaces with water contact angles of around 70◦

[35–37]. These conflicting results may be due to the use of materi-als characterized by different surface topographic parameters such as roughness and micro-texture, and especially different surface chemistry that affect the material wetting behavior. In the present study, the most suitable TiO2-coated substrate, in terms of FN

fib-ril assembly that becomes incorporated into a stable ECM is T9. This sample is characterized by a moderately wettable (water con-tact angle = 55◦) and rough (Ra= 0.71␮m) MOCVD-deposited TiO2

film.

3.4. Assessment of cell proliferation status

To definitively assess the biocompatibility of the native-oxide covered and TiO2-coated Ti substrates, we addressed the

rela-tionship between the properties of analyzed surfaces and cell proliferation. A comparative approach of cell cycle analysis by BrdU incorporation into chromosomal DNA of the S-phase cells (Fig. 5) and PCNA Western blotting assay (Fig. 6) provided valuable information about the proliferation status of HGF-1 cells grown in contact with the specimens.

As shown inFig. 5, the highest density of fluorescent nuclei was observed on the MOCVD-treated T9 sample (Fig. 5b). A lower num-ber of S-phase cells were remarked on the non-treated disc, T1200 (Fig. 5a) with relatively smooth (Ra= 0.35␮m) and moderately

wet-table (contact angle = 60◦) surface. Samples T12 (Fig. 5c) and T14 (Fig. 5d) having highly hydrophilic TiO2deposits and different

sur-face roughness displayed much less labeled cells. Most importantly, Western blot analysis of PCNA demonstrates that the intensity of the band corresponding to PCNA protein (∼36 kDa) decreases in the following order: T9 > T1200 = T12 > T14 (Fig. 6).

Taking together, BrdU immunocytochemistry and Western blot data, we conclude that the proliferation capacity of HGF-1 cells decreases in the following order: T9 > T1200 > T12 > T14. Therefore, the cell proliferation appears to run in parallel with the quality of FN-mediated cell adhesion and is sensitive to surface wettability. Comparison of cell proliferation potential between samples with quasi-similar water contact angle values but of different rough-ness, namely T1200/T9 on one hand and T12/T14 on the other hand, demonstrated that surface roughness could represent a secondary determinant of cell behavior. Thus, HGF-1 cells have proved to pre-fer a moderately rough surface, the optimal Ra value being around 0.70 (the case of T9 and T12 samples). In the literature, cellular adhesion and cell proliferation have been shown to be sensitive to surface roughness[18,38,39]. For example, in their study on the effect of surface topography and composition on cellular response

Fig. 6. Western blotting detection of PCNA expression in lysates of the cells grown in contact with native-oxide covered Ti surface (lane 2), and MOCVD-TiO2coated

specimens: T9 (lane 3), T12 (lane 4), and T14 (lane 5). MW = high molecular weight marker (lane 1).

(8)

Fig. 7. Gelatin-zymographic determination of the MMP secreted by HGF-1 cells maintained in contact for 72 h with: native-oxide covered Ti surface (lane 2), and MOCVD-TiO2coated specimens: T9 (lane 3), T12 (lane 4), and T14 (lane 5). MW-high

molecular weight marker (lane 1).

of human gingival fibroblasts grown on NiTi substrates, Ponsonnet et al.[39]demonstrated that there might exist a roughness thresh-old (between 0.08 and 1␮m) over which cell proliferation became difficult and that the lowest cell proliferation occurred on the NiTi surfaces of the highest roughness.

3.5. MMP gelatinolytic activities expressed in cell culture media It is recognized that cell–biomaterial interface in vitro differs substantially from the periimplant environment in vivo where implant failure can be attributed to several factors such as biolog-ical, microbiological and biomechanical[40]. However, the exact molecular mechanisms of implant failure are not well known. The loosening of implants used in dentistry and orthopedic surgery have been shown to involve proteolytic cascades and MMPs

[41–43]. MMPs are a group of structurally related but geneti-cally distinct endopeptidases that have the capacity to degrade all ECM and basement membrane components[44]. These enzymes are secreted especially by osteoclasts and gingival inflammatory infiltrate and, at high concentrations, they contribute to the resorp-tion/osteolysis and to soft tissue/dental ligament destruction[45]. Considering that interaction with inadequate substrates could induce the cells to synthesize MMPs and eventually detach[28], we evaluated the gelatinolytic activities specific for the MMPs secreted in the conditioned culture media by HGF-1 cells maintained in con-tact with the analyzed specimens for 72 h. As shown inFig. 7, in all analyzed media, the gelatinolytic activities were detected at 72 kDa, 96 kDa and ∼190 kDa corresponding to proMMP-2, proMMP-9 and MMP-dimers respectively. In addition, a barely visible addi-tional gelatinolytic band, corresponding to the active MMP-2 form (68 kDa), was present in each profile. These activities disappeared in the presence of 10 mM 1,10-phenantroline (data not shown). MMP-2 represents a MMP member which is constitutively expressed by many cell types, being associated with physiological tissue remod-elling at low levels of expression, while MMP-9 is mainly involved in inflammatory conditions.

The gelatinolytic activities corresponding to proMMP-9 and especially proMMP-2 were lower for the MOCVD-coated surfaces than for the native-oxide covered surface. These results suggest that in vivo integration of native-oxide covered Ti implants would be more problematic than that of TiO2MOCVD-coated Ti implants.

Taking into account that the intensities of the gelatinolytic bands corresponding to proMMP-2 and proMMP-9 decreased in the fol-lowing order: T14 > T12 > T9, we can suppose that the thickness and crystal size of the TiO2coatings obtained by MOCVD

repre-sent the main surface characteristic of implant material influencing cellular capacity to degrade ECM. In addition, we cannot currently rule out the possibility that other film structural parameters such as roughness, transversal morphology, allotropic composition and wettability contribute to the proteolytic events at material-tissue interface. Therefore, we intend to go deeply into this study and complete it with studies on other MMP members known to be involved in implant failure.

Actually, a long survival of a dental implant is warranted not only by a good osseintegration but also by the presence of an adequate zone of attached gingiva surrounding the implant. The overexpres-sion of MMPs by gingival fibroblasts could be responsible for the reduction in the thickness of the entire gingival soft tissue and of the keratin layer in peri-implant mucosa. As a result of these processes, the products from the oral cavity and peri-implant pocket eas-ily penetrate the peri-implant mucosa. This situation may explain the presence of very high numbers of inflammatory cells around loosened dental implants[46].

The lowest MMP activity expressed in cell culture media by HGF-1 cells grown in contact with T9 recommends this LP-MOCVD prepared thin TiO2 film as being able to induce the best tissues

integration in comparison with the other LP-MOCVD TiO2-coated

Ti samples.

4. Conclusions

In this study, we have confirmed that the LP-MOCVD technique is well adapted to modifying the surface of materials used in den-tal implantology. The present in vitro findings indicate that this technique can provide TiO2covered titanium implants with

supe-rior qualities in terms of FN organization into fibrillar structures, proliferation and host tissue integration capacity. According to our experimental data, the major factors that proved to govern these important biological events for cell–material interactions, are the wettability and, secondarily, the roughness of the oxide surface. The most suitable substrate for the cellular model used in this study is characterized by a moderately rough and wettable TiO2film.

Acknowledgements

Authors are thankful to Yannick Thébault (Institut National Poly-technique de Toulouse) for help in SEM studies and to Bianca Galateanu (University of Bucharest) for expert image analysis.

References

[1] D.F. Williams, Biomaterials 30 (2009) 5897.

[2] M. Geetha, A.K. Singh, R. Asokamani, A.K. Gogia, Prog. Mater. Sci. 54 (2009) 397. [3] M. Textor, C. Sittig, V. Frauchiger, S. Tosatti, D.M. Brunette, Titanium in Medicine: Material Science, Surface Science Biological Responses and Medical Applications, Springer, Heidelberg, Berlin, 2001.

[4] T. Eliades, Int. J. Oral Maxillofac. Implants 12 (1997) 621.

[5] J.C. Keller, C.M. Stanford, J.P. Wightman, R.A. Draughn, R. Zaharias, J. Biomed. Mater. Res. 28 (1994) 939.

[6] D.F. Williams, Biomaterials 29 (2008) 2941.

[7] P.S. Vanzillotta, M.S. Sader, I.N. Bastos, G. Almeida Soares, Dent. Mater. 22 (2006) 275.

[8] L. Le Guéhennec, A. Soueidan, P. Layrolle, Y. Amouriq, Dent. Mater. 23 (2007) 844.

[9] S. Popescu, I. Demetrescu, A.N. Gleizes, Rev. de Chim. 58 (2007) 880. [10] S Popescu, I. Demetrescu, C. Sarantopoulos, A. Gleizes, D. Iordachescu, J. Mater.

Sci.: Mater. Med. 18 (2007) 2075.

[11] S. Popescu, I. Demetrescu, V. Mitran, A.N. Gleizes, Mol. Cryst. Liq. Cryst. 483 (2008) 266.

[12] X. Cui, H.M. Kim, M. Kawashita, L. Wang, T. Xiong, T. Kokubo, T. Nakamura, Dent. Mater. 25 (2009) 80.

[13] K.V. Balla, D.P. DeVasConCellos, W. Xue, S. Bose, A. Bandyopadhyay, Acta Bio-mater. 5 (2009) 1831.

[14] M.P. Casaletto, G.M. Ingo, S. Kaciulis, G. Mattogno, L. Pandol, G. Scavia, Appl. Surf. Sci. 172 (2001) 167.

[15] G. Giavaresi, R. Giardino, L. Ambrosio, G. Battiston, R. Gerbasi, M. Fini, L. Rimon-dini, P. Toricelli, Int. J. Artif. Organs 26 (2003) 774.

[16] G. Giavaresi, L. Ambrosio, G.A. Battiston, U. Casellato, R. Gerbasi, M. Fini, N. Nicoli Aldini, L. Martini, L. Rimondini, R. Giardino, Biomaterials 25 (2004) 5583. [17] K. Mustafa, B. Silva Lopez, K. Hultenby, A. Wennerberg, K. Arvidson, Clin. Oral

Implants Res. 9 (1998) 195.

[18] X. Zhu, J. Chen, L. Scheidelera, R. Reichl, J. Geis-Gerstorfera, Biomaterials 25 (2004) 4087.

[19] D. Kakoli, B. Susmita, B. Amit, Acta Biomater. 3 (2007) 573.

[20] R. Carbonea, I. Marangia, A. Zanardia, L. Giorgettia, E. Chierici, G. Berlanda, A. Podesta, F. Fiorentini, G. Bongiorno, P. Piseri, P.G. Peliccia, P. Milani, Biomaterials 27 (2006) 3221.

(9)

[21] C.M. Advincula, G.F. Rahemtulla, C.R. Advincula, T.E. Adae, E.J. Lemonsa, L.S. Bellisa, Biomaterials 27 (2006) 2201.

[22] A. Ochsenbein, F. Chai, S. Winter, M. Traisnel, J. Breme, F.H. Hildebrand, Acta Biomater. 4 (2008) 1506.

[23] M.M. Bradford, Anal. Biochem. 72 (1976) 248.

[24] A. Cimpean, M. Caloianu, O. Zarnescu, L. Moldovan, N. Efimov, O. Ranga, J. Med. Biochem. 2 (1998) 313.

[25] B. Kasemo, J. Lausmaa, Int. J. Oral Maxillofac. Implants 3 (1988) 83.

[26] C. Sarantopoulos, E. Puzenat, C. Guillard, J.-M. Herrmann, A.N. Gleizes, F. Maury, Appl. Catal. B 91 (2009) 225.

[27] D. Boettiger, J.A. Garcia, Biomaterials 20 (1999) 2427.

[28] G.Rh. Owen, D.O. Meredith, I. Gwynn, R.G. Richards, Eur. Cell. Mater. 9 (2005) 85.

[29] M.S. Frisch, E. Ruoslahti, Curr. Opin. Cell Biol. 9 (1997) 701. [30] R. Pankov, M. Kenneth, J. Cell Sci. 115 (2002) 3861. [31] J. Takagi, Biochem. Soc. Trans. 32 (2004) 403.

[32] G. Altankov, F. Grinnell, T. Groth, J. Biomed. Mater. Res. 30 (1996) 385. [33] T. Groth, G. Altankov, Biomaterials 17 (1996) 1227.

[34] K. Webb, V. Hlady, P.A. Tresco, J. Biomed. Mater. Res. 241 (1998) 422. [35] Y. Ikada, Biomaterials 15 (1994) 725.

[36] J.H. Lee, G. Khang, J.W. Lee, H.B. Lee, J. Colloid Interface Sci. 205 (1998) 323.

[37] N. Faucheux, R. Schweiss, K. Lutzow, C. Werner, T. Groth, Biomaterials 25 (2004) 2721.

[38] L. Ponsonnet, V. Comte, A. Othmane, C. Lagneau, M. Charbonnier, M. Lissac, N. Jaffrezic, Mater. Sci. Eng. C 21 (2002) 157.

[39] L. Ponsonnet, K. Reybier, N. Jaffrezic, V. Comte, C. Lagneau, M. Lissac, C. Martelet, Mater. Sci. Eng. C 23 (2003) 551.

[40] M.C.L.G. Santos, M.I.G. Campos, S.R.P. Line, Braz. J. Oral Sci. 1 (2002) 103. [41] S. Santavirta, T. Sorsa, Y.T. Konttinen, H. Saari, A. Eskola, A.Z. Eisen, Clin. Orthop.

290 (1993) 206.

[42] M. Takagi, S. Santavirta, H. Ida, M. Ishii, Y.T. Konttinen, Clin. Orthop. 352 (1998) 35.

[43] T. Pap, G. Pap, K.M. Hummel, J.K. Franz, E. Jeisy, I. Sainsbury, R.E. Gay, M. Billing-ham, W. Neumann, S. Gay, J. Rheumatol. 26 (1999) 166.

[44] T. Sorsa, L. Tjaderhane, T. Salo, Oral Dis. 10 (2004) 311.

[45] H. Birkedal-Hansen, W.G. Moore, M.K. Bodden, L.J. Windsor, B. Birkedal-Hansen, A. DeCarlo, J.A. Engler, Crit. Rev. Oral Biol. Med. 4 (1993) 197.

[46] B. Liljenberg, F. Gualini, T. Berglundh, M. Tonetti, J. Lindhe, J. Clin. Periodontol. 23 (1996) 1008.

Figure

Fig. 1. Surface micrographs of (a) native-oxide covered Ti surface, and MOCVD-grown TiO 2 films on Ti substrates for different periods of time: (b) 90 min; (c) 250 min, and (d) 360 min; and cross sectional views for the compact (e) and columnar (f) structur
Fig. 2. GIXRD patterns of the TiO 2 films deposited on Ti substrate.
Fig. 4. Indirect immunolocalization of fibronectin adsorbed to and secreted by HGF 1 cells on: (a) native-oxide covered Ti surface, and MOCVD-TiO 2 coated specimens: (b) T9, (c) T12, and (d) T14.
Fig. 5. Immunofluorescent detection of proliferative cells after BrdU incorporation showing the cells that undergo DNA synthesis, as indicated by presence of green bright nuclei, on: (a) native-oxide covered Ti surface, and MOCVD-TiO 2 coated specimens: (b)
+2

Références

Documents relatifs

Our empirical analysis thus far has relied on the implications of the model in identifying the latent factors − output and demand shocks − from the high frequency data on stock,

This allows us to tackle two key issues in the field of third-order nonlinear properties of gold nanoparticles: The influence of the generalized thermal lens in the long laser

The results obtained for both gold and silver nanoparticles show that film morphology strongly depends on laser scanning speed and the number of passages.. We show, furthermore,

There is preliminary experimental evidence for this behavior in high precision LEED measurements (M. Fain, Jr., private communication) of Ar on graphite. At this coverage

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Abstract - Thin oxide films of Ti02, Fe203-and Sn02 have been prepared by the organometallic chemical vapor deposition (MO-CVD) technique.. The structure and character of

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Abstract : ABSTRACT SnO2 thin films were electrodeposited on fluorine tin oxide substrate in nitric acid solution.. The films were found uniform, adherent to the substrate