• Sonuç bulunamadı

Osteogenic activities of polymeric soybean oil-g-polystyrene membranes

N/A
N/A
Protected

Academic year: 2021

Share "Osteogenic activities of polymeric soybean oil-g-polystyrene membranes"

Copied!
18
0
0

Yükleniyor.... (view fulltext now)

Tam metin

(1)

O R I G I N A L P A P E R

Osteogenic activities of polymeric soybean

oil-g-polystyrene membranes

R. Seda Tıg˘lı Aydın•Baki HazerMerve Acar

Menems¸e Gu¨mu¨s¸dereliog˘lu

Received: 30 November 2012 / Revised: 18 February 2013 / Accepted: 30 April 2013 / Published online: 10 May 2013

Ó Springer-Verlag Berlin Heidelberg 2013

Abstract A novel biocompatible copolymer membrane was synthesized and characterized for use in guided bone regeneration using polymeric soybean oil-g-polystyrene (PSO-g-PS) graft copolymer which was successfully obtained by free radical polymerization of styrene initiated by PSO peroxide as a macroinitiator at 80°C. Osteoblastic cellular activities of MC3T3-E1 cells on PSO-g-PS membranes with different soybean oil composition (PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3) were evaluated. Nuclear magnetic resonance (1H NMR) spectra showed that PSO inclusion (mol%) was found to be 27, 69, and 51 % for PSO-g-PS1, PSO-g-PS2, and PS3 membranes, respectively. Superior biocompatibility of the PSO-g-PS membranes was determined compared to polystyrene tissue culture plates (TCPS) as positive control. Cell proliferation was enhanced on PSO-g-PS2 and PSO-g-PS3 membranes compared to PSO-g-PS1 membranes (p \ 0.001), and a statistically significant higher ALP value of MC3T3-E1 cells on PSO-g-PS2 membranes (p \ 0.05) suggested that proliferation and differentiation of preos-teoblastic on PSO-g-PS membranes were enhanced with regard to soybean oil content within the membranes. Thus, the present study suggests that PSO-g-PS2

R. S. Tıg˘lı Aydın (&)

Department of Biomedical Engineering, Bu¨lent Ecevit University, Zonguldak 67100, Turkey e-mail: rseda.tigli@gmail.com

B. Hazer (&)  M. Acar

Department of Chemistry, Bu¨lent Ecevit University, 67100 Zonguldak, Turkey e-mail: bhazer2@yahoo.com; bkhazer@karaelmas.edu.tr

M. Acar

e-mail: merwe802@hotmail.com M. Gu¨mu¨s¸dereliog˘lu

Department of Chemical Engineering, Hacettepe University, 06530 Beytepe/Ankara, Turkey e-mail: menemse@hacettepe.edu.tr

(2)

membranes, which showed a favorable biological environment for the preosteob-lastic cells, can be well suited for bone tissue engineering applications.

Keywords Soybean oil Styrene  Cell–biomaterial interaction  Bone tissue engineering

Introduction

The reconstruction of bone defects, which is a complex process influenced by age, bone structure, vascularization, defect morphology, and adjacent soft tissue [1,2] can be revealed by bone tissue engineering approach. Guided bone regeneration, which is a sub area of tissue engineering, has proven to be a well-established therapy to repair alveolar bone defects affected by periodontal diseases [2,3]. In this therapy, a barrier membrane is placed over bone defects with bone grafting materials to prevent in-growth of fibroblasts and to provide a space for osteogenesis within the underlying blood clot, which encourages bone growth [4,5].

Ideal properties for a barrier membrane should include biocompatibility, space-making effect, osteoinduction, time of absorption compatible with bone regener-ation, and clinical manageability [6–8]. Among diverse biomaterials, currently developed barrier membranes are classified as non-degradable (ePTFE, Gore-TexÒ) [6, 9] and degradable membranes (GuidorÒ, Bio-GideÒ) [6, 10]. Several studies reported that non-degradable membranes have better space-maintaining properties, but a second surgical procedure is required to remove the membranes after new bone generation [11], and another disadvantage is the early removal of membranes with regard to risk of infection [2, 12]. Although degradable membranes are reported as materials which can alter disadvantages of non-degradable membranes [13, 14], there are still challenges with respect to inflammatory reactions in the adjacent tissue, rapid degradation of membranes leading barrier of tissue invasion, and mechanical stability of membrane to sustain surgical treatment [5].

Plant oils are the suitable starting materials for polymers because of their abundance, the rich chemistry that their triglyceride structure provides, and their potential biodegradability [15–25]. Soybean oil, which is one of the polyunsaturated plant oils (MW approx. 874), contains linoleic acid (51 %), oleic acid (25 %), palmitic acid (11 %), linolenic acid (9 %), and stearic acid (4 %) residues [16]. Because of the poly unsaturated inclusion, soya oil is a drying oil and can easily be autoxidized to produce polymeric soya oil peroxides, which behave as a macroperoxy initiator for free radical polymerization of vinyl monomers [26]. Previously, several studies have been reported regarding the synthesis of graft copolymers based on polymeric soya oil peroxide [27–29]. C¸ akmakli et al. [27] reported the auto-oxidation of the unsaturated edible aliphatic oil (soya oil) to prepare polymeric oil peroxy initiators, and they successfully synthesized polymeric soybean oil-graft-methyl methacrylate (PSO-g-PMMA) by peroxidation, epoxida-tion and/or perepoxidaepoxida-tion of soybean oil, and subsequent graft copolymerizaepoxida-tion with MMA. Then, a biodegradability study was performed on PSO-g-PMMA copolymer, which has a non-biodegradable polymer grafted onto a biodegradable

(3)

polymer backbone, and results showed that the copolymer degraded by specific chain end scission [28]. Furthermore, Allı and Hazer [29] demonstrated the polymerization of auto-oxidized polymeric soybean oil with N-isopropyl acrylam-ide. Synthesis of these materials has prompted research into the production of soybean oil biomaterials for tissue regeneration since the effect of soybean oil on fibroblast and macrophage cell adhesion and spreading on polymeric membranes was shown [27]. Furthermore, Hazer et al. [30,31] reported in vivo biocompatibility of the autoxidized and unoxidized unsaturated medium–long chain length (m–lcl) co-poly-3-hydroxyalkanoates (m-lclPHAs) derived from soya oily acids by demonstrating no symptoms such as necrosis, abscess, or tumorigenesis in the vicinity of the implants. Moreover, Liu et al. [32] prepared elastomers with epoxidized soybean oil and epoxidized linseed oil and reported that the elastomers were all biocompatible and completely biodegradable after 3 months in vivo without any granulation or formation of scar tissues. Very recently, Wang et al. [33] reported a new series of soybean-based elastomers poly(epoxidized soybean oil-co-decamethylene diamine), and Miao et al. [34] reported a new class of biocompatible polyurethanes prepared from soybean-oil-based polyol for suitable use in biomedical applications.

In the present study, a novel soybean-oil-based biomaterial, polymeric soybean oil-g-polystyrene (PSO-g-PS) membrane was prepared, characterized, and the cellular activities on the membranes were evaluated with regard to a potential membrane candidate under the scope of guided bone regeneration.

Materials and methods Materials

Soybean oil was locally purchased (C¸ otanak, Turkey) and used as received. Styrene was obtained from Sigma-Aldrich (Germany) and it was purified by vacuum distillation over CaH2. All the other chemicals (Sigma-Aldrich, Germany) were

analytical grade and used without further purification.

Auto-oxidation of soybean oil and the peroxygen analysis

The polymeric soybean oil (PSO) was prepared due the auto-oxidation according to the procedure reported in the literature [27]. In brief, soybean oil (50 g) spread out in a Petri dish (Ø = 16 cm) was exposed to sunlight in the air at room temperature (8 weeks). Then, a gel polymer film associated with a waxy and viscous liquid was formed. Chloroform extraction of the crude polymeric oil for 24 h at room temperature allowed separation of the soluble part of the polymeric soybean oil from the gel. Peroxygen analysis of soluble part fractions was carried out by refluxing a mixture of 2-propanol (50 mL)/acetic acid (10 mL)/saturated aqueous solution of potassium iodide (1 mL) and 0.1 g of the polymeric sample for 10 min and titrating the released iodine against thiosulfate solution.

(4)

Graft copolymerization

Graft copolymers of PSO-g-PS were synthesized using PSO peroxides to initiate styrene monomer during free radical polymerization. In brief, for given amounts of PSO and styrene were charged separately into a Pyrex tube. Argon was introduced through a needle into the tube for about 3 min to expel the air. The tightly capped tube was then put into a water bath at 80°C for 1 day. Then, the contents of the tube were coagulated in methanol. The graft copolymer samples were dried overnight under vacuum at 30°C. Purification of copolymers was assessed using a typical fractional precipitation procedure [35]. Purified copolymers were dried under vacuum at room temperature. PSO-g-PS membranes were prepared by solvent-casting method. For this purpose, 0.5 g of PSO-g-PS was dissolved in 10 mL chloroform and the solution was poured into a glass Petri dish. The solvent was allowed to evaporate overnight at room temperature.

Polymer characterization

Nuclear magnetic resonance (1H NMR) spectrum of PSO and PSO-g-PS was recorded in CDCl3 at 17 °C with a tetramethylsilane internal standard using a

400 MHz/54 mm Ultra Shield Plus NMR (Burker, Ultra long hold time). FTIR analysis of PSO and PSO-g-PS polymers were performed using ATR-FTIR spectrometer (Perkin Elmer Pyris 1) in the range of 400–4,000 cm-1. The molecular weight of the polymeric samples was determined by gel permeation chromatog-raphy (GPC) with a Waters model 6000A solvent delivery system with a model 401 refractive index detector and a mode 730 data module and with two Ultrastyragel linear columns in series. Chloroform was used in the elution at a flow rate of 1.0 mL min-1. A calibration curve was generated with polystyrene standards. In order to evaluate thermal characteristics of PSO-g-PS copolymers, a differential scanning calorimeter (TA Q2000, DSC) was used. The glass transition temperatures (Tg1and Tg3) were evaluated between -100°C and ?140 °C at a heating rate of

10°C/min (first heating) under nitrogen atmosphere, and samples were held at the final temperature for 1 min to eliminate the thermal history applied to the samples. After being cooled to -100°C, they were then reheated to 140 °C at a rate of 10°C/min (second heating). Thermal gravimetric analysis (TGA) of the copolymers was assessed by (TA Q50, TGA) instrument to determine thermal degradation between 0 and 700°C. Biodegradation tests of copolymers were carried out in lipase solution (0.1 mg/mL) in PBS (pH 7.4) containing 0.02 % sodium azide (NaN3). Incubations were performed in a humidified incubator at 37 °C for 14 days.

Dry weights of dehydrated samples were measured at first and second weeks of incubation. Degradation was determined as the percentage of weight loss (WL)

according to Eq. (1).

WLð Þ ¼% ðW0 WÞ=W0

 

 100 ð1Þ

where, W0is initial weight, and W is weight of matrix after degradation.

The wettability of the copolymer samples was assessed by water contact angles measured by sessile drop method at room temperature (Kru¨ss DSA 100, Germany).

(5)

The tensile strength, strain at failure, and elastic modulus of PSO-g-PS membranes were measured for each of the five sheets with a gauge length of 20 mm using a universal testing machine (Zwick Roell, USA) at crosshead speed of 100 mm/min. Cell culture studies

Cell culture studies were carried out with MC3T3-E1, mouse osteogenic cell line (Riken Cell Bank, Ibaraki, Japan). The cells were subcultured in flasks using minimum essential medium (aMEM) (Sigma Co.) supplemented with 10 % (v/v) fetal bovine serum (FBS, Sigma Co.) and 1 % penicillin–streptomycin (Biological Industries, Ashrat, Israel). The cells, maintained at 37°C in a humidified CO2(5 %)

atmosphere (Heraus Instruments, Heraus, Germany), were dissociated with 0.25 % trypsin–ethylenediaminetetraacetic acid (trypsin–EDTA) (Sigma), centrifuged, and resuspended in medium prior to cell seeding. All the cell culture studies were carried out in Laminar Flow Cabinet (Bioair, Type II Laminar Flow Cabinet, Italy). Sterilization and cell seeding

Cell culture studies were conducted in sterile 24-well tissue culture polystyrene (TCPS) dishes in stationary conditions. Prior to cell culture experiments, 24-well TCPS were precoated with 70 % ethanol soaked Parafilm for wrapping the plates, and placed under UV light during 30 min for sterilization. Prepared membranes, which have 15 mm diameter and 100 lm thickness, were sterilized with 70 % ethanol for 1 h, equilibrated in sterile Dulbecco’s PBS (pH 7.4) and then, were immersed in conditioning medium (DMEM-F12 supplemented with 10 % (v/v) FBS) for 2 h prior to cell seeding. Then, 1 mL of cell suspension in osteogenic medium (aMEM ? 10 % FBS containing 50 lg/mL ascorbic acid and 10 mM b-glycerol phosphate) was added to each well to maintain 2 9 104cells mL-1 inoculation density for each membrane. The culture medium was replenished every 2 days.

Cytocompatibility

The cytotoxicity evaluation was performed by direct contact assay. In brief, 1 mL of MC3T3-E1 cell suspension at a density of 2 9 104cells mL-1 was seeded in 24-well plate. Then, sterilized polymer membranes were immersed in culture medium and exposed to MC3T3-E1 cells for 1, 2 and 4 days. As positive control, wells were seeded with MC3T3-E1 cells at the same density in the absence of polymer membranes, and empty wells (without cells) were used as negative controls. Cells on wells were quantitatively assessed with 3-[4, 5-dimethylthiazol-2-yl]-diphenyltetrazolium bromide (MTT) formazan (Sigma, Germany) at different culture periods (MTT assay) at selected time intervals (1st, 2nd, and 4th day). In brief, culture medium was aspirated and washed with 600 lL prewarmed PBS (pH 7.4). Six hundred microlitre prewarmed culture medium supplemented with 60 lL MTT solution (2.5 mg mL-1 MTT dissolved in PBS) was added to each sample, which were incubated at 37°C for 3 h. Then, the medium was removed from each

(6)

well and 400 lL of 0.04 M HCl in isopropanol solution was added to each well to dissolve formazan crystals. The resulting solution was taken out and centrifuged at 13,000 rpm for 2 min. The supernatant was used for measuring optical density spectrophotometrically at 570 nm with reference to 690 nm using a microplate reader (Asys UVM 340, Austria). The data obtained from negative controls were subtracted from measured values. The number of viable cells was correlated to optical density values, and cell viability was then evaluated by normalizing the values to those from positive control wells.

Cell attachment and proliferation

MC3T3-E1 cells were seeded onto polymer membranes and cultured as described above for 2 and 4 h to determine the cell attachment on polymer membranes by crystal violet assay. In brief, the medium was aspirated and cells were washed twice in Dulbecco’s PBS (DPBS). Then, the cells on membranes were fixed with 250 lL of glutaraldehyde (2.5 % v/v) at 4°C for 20 min and stained with crystal violet (0.5 w/v in 20 % methanol) for 30 min. After washing gently with water, the crystal violet was solubilized with 250 lL of 0.5 % sodium dodecyl sulfate (SDS), and the absorbance was measured at 570 nm using a Labomed Double Beam UV–visible spectrophotometer. As a positive control, wells were seeded with MC3T3-E1 cells at the same density in the absence of polymer membranes, and empty polymer membranes (without cells) were used as negative controls. The data obtained from negative controls were subtracted from measured values. The number of attached cells was correlated to optical density values, and cell attachment was then evaluated by determining the ratio of number of cells attached to membranes to the initial cell seeding numbers. MC3T3-E1 cell proliferation on polymer membranes was assessed by MTT assay as described above. Cells seeded on 24-well TCPS were used as control.

Microscopic imaging

After crystal violet staining, cells on polymer membranes were visualized by optical microscope (Olympus, Japan). Cytoskeleton organization on polymer membranes was assessed at the end of 4 h, 1, 2, and 4 days of incubation period. Cells were rinsed twice with PBS, fixed in 2.5 % (v/v) glutaraldehyde in 0.1 M PBS (pH 7.4) for 10 min at 4°C and permeabilized in 0.1 % Triton-X 100 for 5 min. Cell cytoskeletal filamentous actin (F-actin) was visualized by treating the cells with Alexa Fluor 488 phalloidin (Invitrogen, USA) for 20 min, and cell nuclei were counterstained with propidium iodide for 5 min. Samples were visualized using fluorescence microscope (Olympus, Japan).

Alkaline phosphatase (ALP) assay

To evaluate the osteogenic differentiation capacity of the MC3T3-E1 cells, alkaline phosphatase activity (ALP) measurements [36,37] were done at the first week of incubation period. In brief, the membranes were washed thrice with PBS (pH 7.4)

(7)

and freeze-dried at -80°C (Christ, Germany). Then, membranes were cut with scissors and homogenized by sonicating in the presence of 1 % Triton-X 100. The sample lysate was centrifuged at 12,000 rpm for 10 min at 4°C. The supernatant was assayed for ALP activity, using p-nitrophenyl phosphate (pNPP) as substrate at 37°C for 30 min. After stopping reaction with 0.02 N NaOH, hydrolysis of substrate to p-nitro phenol was measured spectrophotometrically at 405 nm using microplate reader (Asys UVM 340, Australia).

Von Kossa analysis

The mineralization of cells on membranes was determined by the von Kossa analysis as described previously [36,38]. In brief, membranes at the second week of culture were removed, washed once with PBS, and then, soaked in ice cold absolute ethanol for 20 min. After washing with deionized water, 5 % AgNO3(0.5 g/mL)

was added onto cells and incubated for 30 min in dark environment. Then, membranes were washed with deionized water and exposed to UV for 2 min. Statistical analysis

All data are expressed as means ± standard deviations. Three similar experiments were done which were carried out in triplicate. Statistical analysis was performed by one-way analysis of variance (ANOVA) in conjunction with Tukey’s post hoc test for multiple comparisons using Graph-Pad Instant (GraphPad Software) statistics program with significance at p \ 0.05. p \ 0.05, p \ 0.01, and p \ 0.001 represent statistically significant, very significant, and extremely significant values, respec-tively, whereas p [ 0.05 represents statistically no significant values.

Results and discussion Auto-oxidation of soybean oil

Polymeric soybean oil containing peroxide/hydroperoxide groups (1.3 wt%) was obtained in the air at room temperature, since polymeric soybean oil peroxide was obtained from auto-oxidized soybean oil, which involves hydrogen abstraction from a methylene group between two double bonds in a polyunsaturated fatty acid chain. Soya oil polymerization reactions via auto-oxidation have been given elsewhere [27]. Molecular weight (Mw) and polydispersity of auto-oxidized PSO were determined from GPC as 9,428 g/g-mole and 2.38, respectively, which are also reported before [27]. Figure1a, b represents FTIR spectra of soybean oil and PSO after auto-oxidation, respectively. As seen from the figure, the characteristic peaks observed at 805 cm-1 (epoxide groups) and 3,300 cm-1 (hydroperoxide groups) clearly indicate auto-oxidation of soybean oil. Figure1c demonstrates 1H NMR spectrum of PSO. The characteristic peaks of the auto-oxidized polymeric soybean oil were marked on the spectrum which confirms the PSO segments in the copolymer structure. Characteristic peaks as indicated in Fig.1c (d, ppm): –CH– of

(8)

soybean oil at 2.8, 2.4, 1.9, 1.4, and 0.9 ppm and the peaks at 4.1–4.4 ppm originate from the protons in the methylene groups of the triglyceride. The vinylic protons are detected at 5.3 ppm.

Graft copolymerization and characterization

The peroxide groups in auto-oxidized soybean oil let PSO to be used as an initiator of styrene during copolymerization. Then, cross-linked and soluble graft copolymer fractions were isolated by means of chloroform extraction. Soluble fractions of the graft copolymers were fractionally precipitated to determine the yield of the graft copolymers. Table1 represents copolymerization conditions and copolymer Fig. 1 FTIR spectra of a soybean oil, b auto-oxidized polymeric soybean oil (PSO). c1H NMR spectrum

of auto-oxidized polymeric soybean oil (PSO). (The characteristic peaks of the auto-oxidized polymeric soybean oil have been marked on the1H NMR spectrum.)

(9)

analysis results as well as molecular weights and polydispersity of the copolymers. Although higher initial amounts of PSO in monomer solution yields higher amount of graft copolymer, optimum polymerization yield was achieved for PSO-g-PS2 copolymer. Figure2a represents FTIR spectra of g-PS1, g-PS2, and PSO-g-PS3 copolymers. The peaks at 1,743 and 1,600 cm-1showed the presence of PSO and polystyrene, respectively. By comparison of the characteristic bands of PSO at 1,743 cm-1 and styrene at 1,600 cm-1 (Fig.2a), PSO amounts (wt%) in the copolymers were calculated using a calibration curve due to the methodology reported before [39], and the results are reported in Table1.1H NMR spectra of copolymer samples are demonstrated in Fig.2b (PSO-g-PS1), Fig.2c (PSO-g-PS2), and Fig.2d (PSO-g-PS3). All samples of PSO-g-PS copolymers contained characteristic peaks as indicated in Fig.2b–d (d, ppm): the aromatic protons of styrene at 6.0–7.6 ppm and the broad signals from 1 to 3 ppm are assignable to CH and CH2protons of the main chain. The characteristic peaks for all copolymers at

2.8, 2.4, 1.9, 1.4, and 0.9 ppm indicate –CH2– of soybean oil in which PSO

inclusion (mol%) was found to be 27, 69 and 51 % (Table1) by taking the ratio of signals at 6.5 ppm (styrene) and 0.9 ppm (PSO) for PSO-g-PS1 (Fig.2b), PSO-g-PS2 (Fig.2c), and PSO-g-PS3 (Fig.2d), respectively.

Thermal properties of the synthesized PSO-g-PS copolymers were determined by DSC and TGA analysis. Glass transition temperatures (Tg) and decomposition

temperatures (Td) of copolymers are listed in Table2. As seen from this table, Tg

values of copolymers were decreased due to the increase of soybean oil content in the copolymer, which may be due to the plasticization effect of PSO in the copolymers that has been observed by lowering the glass transition to 68–84°C compared with 100°C for homo-PS [25]. Thermal decomposition temperatures were shifted to higher values with lower PSO-g-PS decomposition percentages with regard to increase in PSO content within the copolymer (Table2).

Biodegradability of the PSO-g-PS membranes were quantitatively tested, and results indicated that PSO-g-PS2 and PSO-g-PS3 membranes showed gravimetri-cally 6 and 4 % weight loss, respectively; while PSO-g-PS1 and homo-PS membranes had no remarkable weight loss at the end of 14 days of incubation period (Table3). The degree of degradation of the copolymers was related with the Table 1 Results and conditions for the polymerization of styrene initiated by PSO

Copolymer PSO-g-PS PSO (g) Styrene (g) Polymer total (g) Yield soluble wt%

Molecular weight PSO in copolymer Mw 9 104 MWD wta% molb%

PSO-g-PS1 0.2 5.0 3.4 65.4 63.91 3.73 23 27

PSO-g-PS2 2.0 10.0 11.0 91.6 43.08 3.83 68 69

PSO-g-PS3 7.0 30.0 32.0 86.5 32.69 3.82 62 51

a

Calculated from their FTIR spectra (by comparison with peaks at 1,743 cm-1(PSO) and 1,600 cm-1 (styrene), Fig.2)

b Calculated from their NMR spectra (by comparison with peaks at 6.5 ppm (styrene) and 0.9 ppm

(10)

Fig. 2 aFTIR spectra of PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3 copolymers,1H NMR spectra of bPSO-g-PS1, c PSO-g-PS2, and d PSO-g-PS3 copolymers

Table 2 Thermal properties of the PSO-g-PS graft copolymers

Copolymer DSC TGA Tg1(°C) Tg3(°C) Td(°C) PSO-g-PS (wt%) a PSO-g-PS1 -35 84 280 5.9 PSO-g-PS2 -32 68 296 4.9 PSO-g-PS3 -32 75 290 5.4 a Decomposed PSO-g-PS (wt%) at Td

Table 3 Degradation, wettability, and mechanical properties of the graft copolymers Copolymers Degradation (%) Water contact

angle (degree) Tensile strength (MPa) Strain at failure (%) Elastic modulus (MPa) 7th day 14th day PSO-g-PS1 0 0 37.1 ± 5.2 30.7 ± 1.1 1.29 ± 0.08 2033.3 ± 124.7 PSO-g-PS2 2.9 5.9 93.0 ± 1.4 17.9 ± 0.8 1.70 ± 0.17 370.5 ± 82.1 PSO-g-PS3 1.3 4.0 81.6 ± 2.0 20.7 ± 0.7 1.65 ± 0.20 812.5 ± 88.4

(11)

PSO amount in the copolymer, and results clearly concluded that the copolymer PSO-g-PS can undergo enzymatic degradation due to its degradable PSO fragments since PS is not biodegradable. Previously, it has been reported that PSO-g-PMMA copolymer degraded enzymatically through PSO fragments by specific chain end scission, [28] in which our results were in a good correlation. Table3shows water contact angle values of copolymers to evaluate wettability of membranes. According to the results, contact angles of PSO-g-PS membranes were dramatically increased due to the increment of PSO content within the copolymer, which clearly explored that PSO significantly increases hydrophobicity of membranes. The mechanical properties of PSO-g-PS membranes including their tensile strength, elastic modulus, and strain at failure are summarized in Table3. The results showed that, when PSO content was increased within the copolymer, the tensile strength and elastic modulus of the membranes decreased. However, an increase in strain failure due to the increment of PSO content within membranes suggested the membranes with better flexibility with PSO content increment, thus facilitating their manip-ulation in surgical applications.

Cell culture studies

Cell viability, cellular adhesion, and proliferation

Mouse MC3T3-E1 cells were used to evaluate the potentials of PSO-g-PS membranes as a potential candidate of biomaterials for guided bone regeneration. Cell responses to PSO-g-PS membranes were determined with respect to presence of soybean oil within the membranes. No cytotoxicity was shown when MC3T3 cells were cultured in the presence of PSO-g-PS membranes for 1, 2, and 4 days (Fig.3). As seen in Fig.3, the fractions of viable cells, normalized to the positive control (TCPS), were almost 100 % for all copolymers with no statistically significant differences (p [ 0.05) during 4 days of incubation period. Results

(12)

indicated that no leachable toxic molecules were released from the copolymers, leading the good cytocompatibility of membranes.

MC3T3-E1 cell adhesion on copolymer membranes were evaluated and monitored after 2 and 4 h of incubation period (Fig.4). As shown in Fig.4e, j, the presence of soybean oil first decreased the cellular adhesion (Fig.4e) which may be due to the hydrophobic character of soybean oil (Table3). It has been reported that hydrophobic factors can reduce cell adhesion [40], thus the same effect was initially observed. Then, cellular adhesion was sharply increased by the content of soybean oil (Fig.4f). At 4 h, statistically very significant amounts of cellular attachment on PSO-g-PS2 and PSO-g-PS3 membranes was obtained (p \ 0.01) compared to PSO-g-PS1 membranes, indicating supportive effect of soybean oil on cell attachment (Fig.4j). Moreover, the value of cellular attachment on PSO-g-PS2 membranes was almost reached to the value for the positive control (TCPS) with a

Fig. 4 Cellular attachment on PSO-g-PS copolymers. Microscopic images of crystal violet stained MC3T3-E1 cells on a PSO-g-PS1, b PSO-g-PS2, c PSO-g-PS3, and d TCPS after 2 h incubation (magnification 940). Fluorescence microscope images [red stained using propidium iodide, green stained using Alexa Fluor 488-phalloidin) of MC3T3-E1 cells on f PSO-g-PS1, g PSO-g-PS2, h PSO-g-PS3, and iTCPS after 4 h incubation (magnification 940)], the blurred images are caused by membrane opacity. Normalized MC3T3 cell attachment e at 2 h, j at 4 h on PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3 membranes compared with cell-seeded TCPS as positive control. Statistical differences between groups (n = 3, TCPS is control group, ***p \ 0.001, *p \ 0.05; PSO-g-PS1 is control group,xxxp\ 0.001)

(13)

small statistical difference (p \ 0.05). Similar supportive results with regard to soybean oil on cell viability and cell attachment were reported which confirms our results [32]. Figure 4a–d, f–i demonstrates the morphological character of attached MC3T3-E1 cells on copolymeric membranes and TCPS. Unlike at 2 h (Fig.4a–d), after 4 h incubation, cells on all types of membranes started to get in touch with each other with a similar cell morphology compared to TCPS (Fig.4f–i).

Proliferation of MC3T3-E1 cells on PSO-g-PS copolymers were evaluated and assessed by MTT assay (Fig.5b). As demonstrated in Fig.5b, for the first 2 days, metabolic activities of MC3T3-E1 cells on PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3 membranes were lower compared to TCPS (p \ 0.001). Moreover, for the first 2 days, there is no statistically significant difference among the metabolic activities of MC3T3-E1 cells on PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3 membranes (p [ 0.05) and PSO-g-PS2 and PSO-g-PS3 membranes (p [ 0.05). TCPS is generally evaluated as a control and should be interpreted as demonstrating the maximum proliferation ability. However, at the end of 4 days, a sharp increase of optical density (OD) values was achieved especially for PSO-g-PS2 and PSO-g-PS3 membranes, while OD values of TCPS was decreased (Fig.5b) which is a real effect as cells become crowded and make a confluent layer through TCPS (after 4 days). As seen from Fig.5b, cell proliferation was enhanced on PSO-g-PS2 and PSO-g-PS3 membranes compared to PSO-g-PS1 membranes (p \ 0.001) and suggesting the superior effect of soybean oil on osteoblastic cell proliferation. Lukyanova et al. [41] reported the effect of soybean-oil-based scaffold on Chinese hamster ovary (CHO) fibroblast cell growth, and results suggested that soybean oil induces an environment which favors cell migration.

Figure5a represents MC3T3-E1 cells grown on PS1, PS2, PSO-g-PS3 membranes, and TCPS after 1, 2, and 4 days of culture period. As shown in Fig.5a, cellular morphologies of MC3T3-E1 cells on PSO-g-PS membranes were dramatically changed compared to TCPS, especially on days 1 and 2. Cells on PSO-g-PS membranes tend to get contact with other cells in a spindle-like morphology compared to cells grown on TCPS (Fig.5a). Moreover, there were visible thin actin stress fibers with prominent focal contacts which may lead to important consequences for cell processes such as migration and proliferation [42]. Cytoskeleton organization in the fully spread MC3T3-E1 cells on PSO-g-PS membranes showed enhanced cell migration and cell proliferation at the end of 4 days of incubation.

ALP activity and mineral deposition

Investigation and identification of cellular activities on a biomedical material provide useful data to develop new therapeutic applications [43, 44]. For bone regeneration, the activities of ALP and deposit minerals in vitro are dependent to the synthesis of ECM [36,37]. MC3T3-E1 cells are osteogenic cell line and have the capacity to differentiate into osteoblasts and deposit minerals in vitro [45]. ALP activity, which is the marker associated with ECM development, has been routinely used as an early marker of osteoblasts differentiation [42]. ALP activity of MC3T3-E1 cells and mineral deposition on PSO-g-PS membranes was characterized to

(14)

evaluate osteoblastic differentiation (Fig.6). As demonstrated in Fig.6a, ALP activities of MC3T3-E1 cells on PSO-g-PS2 membranes showed the greatest statistically different value (p \ 0.05) compared to all. Moreover, a statistically significant higher ALP value of MC3T3-E1 cells on PSO-g-PS2 membranes compared to PSO-g-PS3 membranes (p \ 0.05) were obtained, suggesting the effect of soybean oil on ALP activity (Fig.6a).

Fig. 5 aFluorescent microscope images (red stained using propidium iodide, blue stained using Alexa Fluor 488-phalloidin) of MC3T3-E1 cells on PSO-g-PS1, PSO-g-PS2, PSO-g-PS3 membranes, and TCPS at 1st, 2nd, and 4th days of incubation (magnification 940), the blurred images are caused by membrane opacity. b MTT assay of PSO-g-PS1, PSO-g-PS2, PSO-g-PS3 membranes, and TCPS. Statistical differences between groups (n = 3, ***p \ 0.001; TCPS is control group,xxxp\ 0.001; PSO-g-PS1 is

(15)

Von Kossa staining was used to visualize mineralized nodule formation on cell-seeded membranes. Different levels of staining (black-brown areas) were observed in each group (Fig.6b). As shown in Fig.6b, mineral deposition (black-brown areas) was maximized at PSO-g-PS2 membranes. Results concluded that soybean-oil-based membranes not only increased cell proliferation, but also enhanced differentiation of osteoblasts. Taking all these data into account, PSO-g-PS2 membranes showed a favorable biological environment which induces functions of the preosteoblastic cells. All these information may provide a scientific basis for the use of PSO-g-PS membranes as a barrier membranes used in bone and periodontal tissue engineering applications.

Conclusion

In this study, a novel barrier membrane based on soybean oil was investigated in terms of improving our understanding on the roles of different material character-istics in guided tissue regeneration. PSO-g-PS membranes were successfully prepared using auto-oxidized soybean oil as a macroperoxy initiator for free radical polymerization of styrene. Cellular behavior of MC3T3-E1 cells on membranes Fig. 6 aALP activity of MC3T3-E1 cells cultured on PSO-g-PS1, PSO-g-PS2, PSO-g-PS3 membranes, and TCPS after 1 week of incubation. Statistical differences between groups (n = 3, *p \ 0.05; TCPS is control group). b von Kossa stained PSO-g-PS1, PSO-g-PS2, and PSO-g-PS3 membranes after 2 weeks of incubation. Black-brown areas indicate mineral deposition on membranes

(16)

were evaluated in terms of cell attachment, proliferation, and differentiation. Results concluded that PSO-g-PS2 (the highest molar PSO inclusion in copolymers; 69 %) membranes were a good candidate of a supportive membrane due to their cytocompatibility and in vitro degradability for use in biomedical applications. However, further in vivo studies should be performed to highlight in vivo absorbable characteristics of the membranes which is a very challenging area concerning the development of guided bone regeneration membranes.

Acknowledgments The Authors thank to TU¨ BI˙TAK (Grant No. 211T016) and Bu¨lent Ecevit University (Grant# 2011-10-3-02) for financial support.

References

1. Jung RE, Zwahlen R, Weber FE, Molenberg A, Van Lenthe GH, Hammerle CH (2006) Evaluation of an in situ formed synthetic hydrogel as a biodegradable membrane for guided bone regeneration. Clin Oral Implants Res 17:426–433

2. Ha¨mmerle CH, Jung RE (2003) Bone augmentation by means of barrier membranes. Periodontology 2000(33):36–53

3. Hallman M, Sennerby L, Lundgren S (2002) A clinical and histologic evaluation of implant inte-gration in the posterior maxilla after sinus floor augmentation with autogenous bone, bovine hydroxyapatite, or a 20:80 mixture. Int J Oral Maxillofac Implants 17(5):635–643

4. De Macedo NL, De Macedo LG, Monteiro Ado S (2008) Calcium sulfate and PTFE nonporous barrier for regeneration of experimental bone defects. Med Oral Patol Oral Cir Bucal 13:E375–E379 5. Fujihara K, Kotaki M, Ramakrishna S (2005) Guided bone regeneration membrane made of

polycaprolactone/calcium carbonate composite nano-fibers. Biomaterials 26:4139–4147

6. Gielkens PFM, Schortinghuis J, de Jong JR, Raghoebar GM, Stegenga B, Bos RRM (2008) Vivosorb (R), Bio-gide (R), and Gore-tex (R) as barrier membranes in rat mandibular defects: an evaluation by microradiography and micro-CT. Clin Oral Implants Res 19:516–521

7. van Leeuwen AC, Huddleston Slater JJ, Gielkens PF, de Jong JR, Grijpma DW, Bos RR (2012) Guided bone regeneration in rat mandibular defects using resorbable poly(trimethylene carbonate) barrier membranes. Acta Biomater 8:1422–1429

8. Monteiro ASF, Macedo LGS, Macedo NL, Balducci I (2010) Polyurethane and PTFE membranes for guided bone regeneration: histopathological and ultrastructural evaluation. Med Oral Patol Oral Cir Bucal 1(15):e401–e406

9. Piattelli A, Scarano A, Paolantonio M (1996) Bone formation inside the material interstices of e-PTFE membranes: a light microscopical and histochemical study in man. Biomaterials 17:1725–1731 10. Piattelli A, Scarano A, Russo P, Matarasso S (1996) Evaluation of guided bone regeneration in rabbit

tibia using bioresorbable and non-bioresorbable membranes. Biomaterials 17:791–796

11. Stavropoulos F, Dahlin C, Ruskin JD, Johansson C (2004) A comparative study of barrier membranes as graft protectors in the treatment of localized bone defects. An experimental study in a canine model. Clin Oral Implan Res 15:435–442

12. Machtei EE (2001) The effect of membrane exposure on the outcome of regenerative procedures in humans: a meta-analysis. J Periodontol 72:512–516

13. Huang J, He H, Sheng L, Gu G (2002) Comparison of calcium alginate film with collagen membrane for guided bone regeneration in mandibular defects in rabbits. J Oral Maxillofac Surg 60:1449–1454 14. Kikuchi M, Koyama Y, Yamada T, Imamura Y, Okada T, Shirahama N, Akita K, Takakuda K, Tanaka J (2004) Development of guided bone regeneration membrane composed of b-tricalcium phosphate and poly (L-lactide-co-glycolide-co-e-caprolactone) composites. Biomaterials 25:5979–5986

15. Can E, Wool RP, Ku¨sefog˘lu S (2006) Soybean and castor oil based monomers, synthesis and copolymerization with styrene. J Appl Polym Sci 102:2433–2477

16. Ilter S, Hazer B, Borcakli M, Atici O (2001) Graft copolymerization of methyl methacrylate onto bacterial polyester containing unsaturated side chains. Macromol Chem Phys 202:2281–2286 17. Biermann U, Bornscheuer U, Meier MAR, Metzger JO, Schafer HJ (2011) Oils and fats as renewable

(17)

18. Guner FS, Usta S, Erciyes AT, Yagci Y (2000) Styrenation of triglyceride oils by macromonomer technique. J Coat Technol 72:107–110

19. Ligadas G, Ronda JC, Galia M, Cadiz V (2010) Plant oils as platform chemicals for polyurethane synthesis: current state-of-the-art. Biomacromolecules 11:2825–2835

20. Tan SG, Chow WS (2011) Thermal properties, curing characteristics and water absorption of soybean oil-based thermoset. Express Polym Lett 5:480–492

21. Soucek MD, Khattab T, Wu J (2012) Review of autoxidation and driers. Prog Org Coat 73:435–454 22. Acar M, C¸ oban S, Hazer B (2013) Novel water soluble soya oil polymer from oxidized soya oil

polymer and diethanolamine. J Macromol Sci Part A: Pure Appl Chem 50:287–296

23. Narine SS, Kong X, Bouzidi L, Sporns P (2007) Physical properties of polyurethanes produced from polyols from seed oils: II foams. J Amer Oil Chem Soc 84:65–72

24. Xia Y, Larock RC (2010) Vegetable oil-based polymeric materials: synthesis, properties, and applications. Green Chem 12:1893–1909

25. Cakmakli B, Hazer B, Tekin IO, Kizgut S, Koksal M, Menceloglu Y (2004) Synthesis and char-acterization of polymeric linseed oil grafted methyl methacrylate or styrene. Macromol Biosci 4:649–655

26. Keles¸ E, Hazer B (2008) Autooxidized polyunsaturated oils/oily acids: post-it applications and reactions with Fe(III) and adhesion properties. Macromol Symp 269:154–160

27. C¸ akmakli B, Hazer B, Tekin I˙O¨ , Co¨mert FB (2005) Synthesis and characterization of polymeric soybean oil-g-methyl methacrylate (and n-butyl methacrylate) graft copolymers: biocompatibility and bacterial adhesion. Biomacromolecules 6:1750–1758

28. Vinu R, Marimuthu A, Giridhar M (2010) Enzymatic degradation of poly(soybean oil-g-methyl methacrylate). J Polym Eng 30:57–76

29. Allı A, Hazer B (2008) Poly(N-isopropyl acrylamide) thermoresponsive cross-linked conjugates containing polymeric soybean oil and/or polypropylene glycol. Eur Poly J 44:1701–1713 30. Hazer DB, Kilicay E, Hazer B (2012) Poly (3-hydroxyalkanoate)s: diversification and biomedical

applications. A state of the art review. Mat Sci Eng C 32:637–647

31. Hazer DB, Hazer B, Kaymaz F (2009) Synthesis of microbial elastomers based on soybean oily acids. Biocompatibility studies. Biomed Mater 4:035011

32. Liu Z, Xu Y, Cao L, Bao C, Sun H, Wang L, Daib K, Zhu L (2012) Phosphoester cross-linked vegetable oil to construct a biodegradable and biocompatible elastomer. Soft Matter 8:5888–5895 33. Wang Z, Zhang X, Wang R, Kang H, Qiao B, Ma J, Zhang L, Wang H (2012) Synthesis and

characterization of novel soybean-oil-based elastomers with favorable processability and tunable properties. Macromolecules 45(22):9010–9019

34. Miao S, Sun L, Wang P, Liu R, Su Z, Zhang S (2012) Soybean oil-based polyurethane networks as candidate biomaterials: synthesis and biocompatibility. Eur J Lipid Sci and Technol 114:1165–1174 35. Wu B, Lenz RW, Hazer B (1999) Polymerization of methyl methacrylate and its copolymerization

with caprolactone catalyzed by isobutyl alumoxane catalyst. Macromolecules 32:6856–6859 36. Akman AC, Tıg˘lı RS, Gu¨mu¨s¸dereliog˘lu M, Nohutc¸u RM (2010) Bone morphogenetic

protein-6-loaded chitosan scaffolds enhance the osteoblastic characteristics of MC3T3-E1 cells. Artif Organs 34(1):65–74

37. Gu¨mu¨s¸dereliog˘lu M, Karakec¸ili A, Demirtas¸ TT (2011) Osteogenic activities of MC3T3-E1 cells on heparin-immobilized poly(caprolactone) membranes. J Bioact Compat Pol 26(3):257–267 38. Akman AC, Tıg˘lı RS, Gu¨mu¨s¸dereliog˘lu M, Nohutc¸u RM (2010) bFGF-loaded HA–chitosan: a

promising scaffold for periodontal tissue engineering. J Biomed Mater Res A 92A(3):953–962 39. Hazer B (1995) Grafting on polybutadiene with macro or macromonomer initiators containing

poly(ethylene glycol) units. Macromol Chem Phys 196:1945–1952

40. Ajiro H, Akashi M (2010) Cell proliferation on stereoregular isotactic-poly(propylene oxide) as a bulk substrate. Biomacromolecules 11:2840–2844

41. Lukyanova L, Franceschi-Messant S, Vicendo P, Pereza E, Rico-Lattes I, Weinkamer R (2010) Preparation and evaluation of microporous organogel scaffolds for cell viability and proliferation. Colloid Surf B 79:105–112

42. Li J, Yun H, Gong Y, Zhao N, Zhang X (2006) Investigation of MC3T3-E1 cell behavior on the surface of GRGDS-coupled chitosan. Biomacromolecules 7:1112–1123

43. Tıg˘lı RS, Kazarog˘lu NM, Mavis¸ B, Gu¨mu¨s¸dereliog˘lu M (2011) Cellular behavior on epidermal growth factor (EGF)-immobilized PCL/gelatin nanofibrous scaffolds. J Biomater Sci Polymer Ed 22:207–223

(18)

44. Tıg˘lı RS, Gu¨mu¨s¸dereliog˘lu M (2009) Chondrogenesis on BMP-6 loaded chitosan scaffolds in sta-tionary and dynamic cultures. Biotechnol Bioeng 104(3):601–610

45. Sudo H, Kodama HA, Amagai Y, Yamamoto S, Kasai S (1983) In vitro differentiation and calci-fication in a new clonal osteogenic cell line derived from newborn mouse calvaria. J Cell Biol 96:191–198

Referanslar

Benzer Belgeler

[r]

Güler, Mehmet ve

 (2) Changes in the membrane potential can cause movement of the charged regions on a channel protein —voltage-gated channel.  (3) Stretching the membrane may affect the

 When solutions of different osmolarities are separated by a membrane permeable to water, but not to solute, water will move from the side with lower osmolarity to the side

 In addition to having a binding site for the actively transported solute, the transport protein also has a binding site for an ion - This ion is usually sodium.  the energy

There are, however, substantial differences among species in the process of implantation, particularly with regard to "invasiveness," or how much the embryo erodes

• Increase in cytosolic Ca2+ concentration activates proteins required for the vesicle membrane to fuse with the plasma membrane and release the vesicle contents into

It turns out that for a general domain G in the complex plane there is always a SOT continuous up to the boundary of harmonic mapping on G for a given SOT continuous function on