Toughening and strengthening mechanisms of porous akermanite scaffolds reinforced with nano-titania

Pei Fengb, Chengde Gaob, Cijun Shuai*bc and Shuping Peng*ad
aHunan Provincial Tumor Hospital and the Affiliated Tumor Hospital of Xiangya School of Medicine, Central South University, Changsha, 410013, China. E-mail: shuping@csu.edu.cn
bState Key Laboratory of High Performance Complex Manufacturing, Central South University, Changsha, 410083, China. E-mail: shuai@csu.edu.cn; Fax: +86-731-88879044; Tel: +86-731-88879351
cOrthopedic Biomedical Materials Institute, Central South University, 410083, China
dSchool of Basic Medical Science, Central South University, Changsha, 410078, China

Received 10th October 2014 , Accepted 5th December 2014

First published on 5th December 2014


Abstract

Akermanite possesses excellent biocompatibility and biodegradability, while low fracture toughness and brittleness have limited its use in load bearing sites of bone tissue. In this work, nano-titania (nano-TiO2) was dispersed into the ceramic-matrix to enhance the mechanical properties of porous akermanite scaffolds fabricated with selective laser sintering (SLS). The fabrication process, microstructure and mechanical and biological properties were investigated. The results showed that the nano-TiO2 particles were dispersed both within the akermanite grains and along the grain boundaries. The grain size of akermanite was refined due to the pinning effect of the nano-TiO2 particles on the grain boundaries. The crack deflection around the nano-TiO2 particles was observed due to the mismatch of thermal expansion coefficients between TiO2 and akermanite. The fracture mode changed from intergranular fracture to more and more transgranular fracture as the concentration of nano-TiO2 increased from 0 to 5 wt%. Meanwhile, the fracture toughness, Vickers hardness, compressive strength and stiffness were significantly increased with increasing nano-TiO2. The improvement of mechanical properties was due to the grain size refinement, the crack deflection, as well as the fracture mode transition. The bone like apatite was formed on the scaffolds in simulated body fluid (SBF). The human osteoblast-like MG-63 cells (MG-63 cells) adhered and grew well on the scaffolds. The porous akermanite scaffolds reinforced with nano-TiO2 have considerable potential for application in bone tissue engineering.


1 Introduction

Bone tissue engineering has a great potential for repairing bone defects.1 Scaffolds play a vital role in bone tissue engineering as they provide a framework and initial support for cell attachment, proliferation and differentiation. An ideal scaffold should possess good biocompatibility, adequate mechanical properties and controllable degradation rates.2 Bioceramics are one of the most potential scaffold materials due to the biocompatibility and biodegradability. Akermanite [Ca2MgSi2O7], as a Ca-, Si-, Mg-containing bioceramic, has been demonstrated to be osteoinductive and able to promote bone repair.3,4 It can release Ca, Si and Mg ions after implantation in the body. These ions can enhance osteoblasts adhesion, proliferation and differentiation.5,6 Moreover, akermanite has a controllable degradation rate and can form a bone like apatite layer after immersion in simulated body fluid.7,8 The preliminary studies indicated that akermanite has great potential as a scaffold material. However, the mechanical properties are much lower than the human cortical bone, which limits the application in heavy loaded implants.9–11 Therefore, it is necessary to further increase the mechanical properties of akermanite.

The mechanical properties of ceramics are known to be improved significantly by dispersing nanometer sized particles such as metal oxides,12–14 carbon nanotube,15–17 bioglass (BG)17–20 into the ceramic-matrix. Among the many reinforcements used in previous studies, TiO2 have attracted considerable attention mainly due to the excellent biocompatibility and high mechanical characteristics.21,22 Khalil et al.23 used TiO2 as reinforcement to improve the mechanical properties of hydroxyapatite (HAP) and found that the addition of 5 wt% TiO2 could improve the fracture toughness and compressive strength of sintered compacts. Kailasanathan et al.24 prepared nano-HAP/gelatin composites with incorporation of TiO2 ceramics as reinforcing phase and found that the compressive strength of the TiO2 enhanced scaffolds increased more than twice than nano-HAP/gelatin composites. Seeley et al.25 fabricated the tricalcium phosphate (TCP) compacts with different TiO2 content and found that the presence of TiO2 in TCP improved densification and compression strength. Despite these researchers have used TiO2 to improve the mechanical properties of bioceramics, the details of the mechanisms leading to toughening and strengthening was still unclear.

In addition, Scaffolds should possess an appropriate porous structure (pore size, porosity, pore distribution and geometry shape),26–28 which is critical to offer good conditions for cells growth and flow transport of nutrients and metabolic waste after the implantation. The technique of SLS can be used to manufacture a 3D porous scaffold with customized external shape.29,30 Moreover, it can realize the accurate control of porous structure to meet the spatial structure requirements of bone tissue regeneration.31,32 In the present work, the interconnected porous akermanite scaffolds reinforced with nano-TiO2 were successfully fabricated by SLS. The fracture toughness, Vickers hardness, compressive strength and stiffness were measured as a function of nano-TiO2. The effects of nano-TiO2 on the mechanical properties were investigated and discussed. The toughening and strengthening mechanisms of nano-TiO2 reinforced porous akermanite scaffolds were analyzed. In addition, the bioactivity and cytocompatibility of the akermanite scaffolds were evaluated.

2 Materials and methods

2.1 Materials

The starting akermanite powder (purity ≥ 98%) was obtained from Kunshan Chinese Technology New Materials Co. Ltd. The akermanite powder was synthesized by a sol–gel method using tetraethyl silicate (C8H20O4Si), magnesium nitrate (Mg(NO3)2·6H2O) and calcium nitrate (Ca(NO3)2·4H2O). The starting nano-TiO2 powder (rutile) (purity ≥ 99.9%) was acquired from Nanjing Emperor Nano Material Co. Ltd. Proportions with weight fractions ranging from 0 to 8 wt% of nano-TiO2 powder were added to the starting akermanite powder. The mixture was firstly milled in an agate mortar for 1 h, and then appropriate amounts of ethanol were added followed by ball milling for 8 h. The SEM micrographs of akermanite with different contents of nano-TiO2 after mechanical mixing were shown in Fig. 1. The nano-TiO2 particles were uniform distribution in the akermanite matrix when the contents of nano-TiO2 were 1 wt% and 5 wt% (Fig. 1a and b). The nano-TiO2 particles did not disperse well with a little agglomeration when the content of nano-TiO2 was 8 wt% (Fig. 1c). Previous studies have shown that the agglomeration of nano-particles led to a significant decrease in mechanical properties.33–35 Therefore, in this paper, the akermanite powder was mixed with different proportions of 0 wt%, 1 wt%, 2 wt%, 3 wt%, 4 wt%, 5 wt%, 6 wt% and 7 wt% of nano-TiO2 powder to form eight akermanite based composite powder.
image file: c4ra12095g-f1.tif
Fig. 1 SEM micrographs of the akermanite/nano-TiO2 after mixing (a) akermanite + 1 wt% nano-TiO2, (b) akermanite + 5 wt% nano-TiO2, (c) akermanite + 8 wt% nano-TiO2.

2.2 Scaffold design and fabrication

A porous scaffold model (L × W × H = 14 × 14 × 5 mm3) with the 3D interconnected pore structure was designed using SolidWorks software (Solidworks 2009, Concord, MA), as shown in Fig. 2a and b. When a porous scaffold was fabricated, the design model was exported in the STL format and then the computer sliced the model into a stack of thin slices. The laser sintering experimental was carried out on a homemade SLS system (a detail description of the sintering system and the parameters are explained in the reference).36 Firstly, a ceramic substrate was placed on the sintering platform. And then, a uniform thin layer of powder was deposited on the substrate, with the layer thickness of 0.1–0.2 mm. Laser was guided to sinter the powder selectively to form a layer. The following processing parameters were used: spot size of 1.2 mm, laser power of 7.8 W, laser scanning speed of 100 mm min−1, and scan line spacing of 2.7 mm. The motion platform changes the sintering direction to y-axis after the completion of the sintering for a strut along x-axis. The similar process was repeated many times until the scaffold was completed. Last, the excess powder surrounding the scaffold was brushed off. The 3D porous scaffold is shown in Fig. 2c and d. It can be seen that the scaffold has an open, uniform and interconnected porous structure.
image file: c4ra12095g-f2.tif
Fig. 2 (a) Top plane view and (b) cross-section view of the 3D structures of scaffold model: d1 = 2.7 mm; d2 = 1.5 mm; d3 = 14.7 mm; d4 = 5 mm, (c) top plane view and (d) cross-section view of the fabricated porous scaffold.

2.3 Characterization

The starting powder, the 3D porous scaffold, the Vickers hardness indentation, the grains and their boundaries and the crack were observed by scanning electron microscopy (SEM, JSM-5600LV, Japan). The phase compositions of the starting powders, the composite powders, the fabricated scaffolds and the surface deposition after soaking in SBF were determined by X-ray diffraction (XRD, D8-ADVANCE Bruker AXS Inc., Germany), using Cu-Kα radiation at 40 mA and 40 kV. Data were recorded in the 2θ range of 20–50° (step size 0.02° and scanning rate 8° min−1). The obtained experimental patterns were compared to the standards joint committee on powder diffraction and standards (JCDPS), which involved card # 09-0432 for HAP, # 87-0052 for akermanite and # 76-0317 for rutile TiO2. Scaffold specimens were polished with diamond pastes down to 1.5 μm and subsequently etched by immersing in a 2 wt% hydrofluoric acid (HF) solution for 30 s to delineate the grain boundary. Each specimen was thoroughly washed with distilled water and rinsed to remove any residual acid after etching. The specimens were dried at room temperature, gold coated, and then examined under high vacuum. The chemical composition of the starting powder and the specimens after etching was evaluated by energy dispersive X-ray spectroscopy (EDS) analysis together with SEM.

Both Vickers hardness and fracture toughness were measured with a Vickers Microindenter (Digital Micro Hardness Tester, HXD-1000 TM/LCD, Shanghai Taiming Optical Instrument Co. Ltd). Prior to indentation, the struts specimens (1 × 1.2 × 1.2 mm3) were mounted in a resin and their surfaces were ground and polished. The indentation tests were performed at a load of 300 gf and a hold of 15 s. The Vickers hardness and fracture toughness were calculated from eqn (1) (ref. 37) and (2) (ref. 38), respectively.

 
HV = 0.1891F/d2 (1)
 
KIC = 0.0824F/C3/2 (2)
where HV is the Vickers hardness (GPa), KIC is the fracture toughness (MPa m1/2), F is the applied load (MN), d is the average diagonal line length of the indentation (m), and C the radial crack length (m) measured from the center point of the indentation impression.

The compressive strength and stiffness of the scaffolds (14 × 14 × 5 mm3) were carried out using microcomputer control electronic universal test machine (WD-D1, Shanghai zhuoji instruments Co. Ltd). The crosshead speed was set at 0.5 mm min−1, and the load was axially applied until the scaffold was failed. During the compressive test, the load and displacement were monitored and recorded. Five specimens were prepared from individual compositions for each type of mechanical properties test. Then the average value was taken and the maximum error obtained was found to be less than 5%.

2.4 Apatite-formation ability of the scaffolds in SBF

The bioactivity of the nano-TiO2 reinforced akermanite scaffolds (14 × 14 × 5 mm3) was investigated by immersion in SBF solutions at 37 °C for 2 and 6 days with the following ionic concentrations (in units of mM): 142.0 Na+, 5.0 K+, 1.5 Mg2+, 2.5 Ca2+, 147.8 Cl, 4.2 HCO3, 1.0 HPO42− and 0.5 SO42−. The solutions were buffered at pH = 7.40 with tris-hydroxymethyl-amminomethane [(CH2OH)3CNH2] and hydrochloric acid [HCl] according to the procedure described by Kokubo et al.39 After the set soaking time, the scaffolds were removed from SBF, gently rinsed with distilled water, and then dried at room temperature in a drying oven for 5 h before further characterization. The surfaces morphology of the scaffolds was observed using SEM and the formation of bone like apatite on the surface of the scaffolds was characterized by EDS.

2.5 Cell attachment and morphology of MG-63 on the scaffolds

The human osteoblast-like MG-63 cells were obtained from American Type Culture Collection (Rockville, MD, USA) and cultured as previously described.40 The proliferation medium contained 500 ml MG-63 cells basal medium (DMEM; Cell-gro by Mediatech, Inc., Manassas, VA) supplemented with 10% fetal bovine serum (Gibco, Carlsbad, CA) and 1% penicillin/streptomycin (Gibco). Cells were cultured in a humidified 37 °C/5% CO2 incubator and the culture medium was changed every 2 days. The morphology of primary MG-63 cells was observed using a light microscope (Olympus IX51, Japan). 4 × 105 cells were added to the porous scaffolds (14 × 14 × 5 mm3) in 96 well plates.

For the measurement of cell adhesion, the number of living cells was measured according to 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide (MTT) assays. After 3, 7, and 10 days cell incubation on the pure akermanite scaffold and the 5 wt% nano-TiO2 reinforced akermanite scaffold, non-adherent cells were removed by twice gentle washes in PBS. About 1 ml of 16% MTT in DMEM medium was added to each well and incubated at 37 °C in a 5% CO2 incubator for 4 h. Then 100 μl of the homogenized solution was transferred to the 96 well plates and the optical density was recorded at 570 nm using an enzyme immunosorbent assay reader. By preparing a standard curve for MTT reduction against the cell number, optical-density values can be converted into actual cell numbers. For morphological observation, the scaffolds were removed form the culture wells, rinsed in PBS, and then fixed with 2.5% glutaraldehyde in PBS for 1 h. Subsequently, the scaffolds were dehydrated through a series of graded ethanol alcohols and dried with hexamethyldisilazane at room temperature. The scaffolds were coated with a very thin layer of gold and observed by SEM.

3 Results and discussion

3.1 Porous akermanite scaffolds

The overview of the scaffolds has a well ordered and interconnected pore channel structure (Fig. 3a). The actual size of the struts is 1.2 ± 0.12 mm, while the designed dimension of the struts is 1.2 mm. The reason for this little offset is because some powders in the design pore area are partially fused together. The pore size of the scaffold is around 1.5 mm (Fig. 3b). It has been reported that the better osteogenesis for implants when using pores greater than 300 μm, due to enhanced new bone formation and the formation of capillaries.41 Hollister et al.42 have carried out in vivo studies on hydroxyapatite scaffolds with pore diameters ranging between 400 and 1200 μm in a minipig mandibular defect model. They found significant bone growth on designed scaffolds for all pores, with no statistical difference between pore sizes. Roy et al.43 also found that there was no significant difference in bone growth for 500 and 1600 μm pores for poly (DL-lactic-co-glycolic acid) (PLGA) scaffolds made by a 3D printing technique. There is good bonding between the adjacent layers of the strut (Fig. 3c). This structural characteristic is desirable for improving the mechanical properties of the scaffold. The strut has a very smooth and dense surface (Fig. 3d). It can be inferred that SLS technique can be used to fabricate the scaffold with 3D interconnected pore structures and the laser energy is enough to melt the composite powder with the predetermined process parameters.
image file: c4ra12095g-f3.tif
Fig. 3 (a) Overview of the pore channels, (b) a single pore channel, (c) the staggered layer structure, (d) the strut surface.

The XRD pattern of the starting powders, the akermanite scaffold with and without TiO2 was shown in Fig. 4. There were presented the characteristic peaks of akermanite and rutile TiO2, and no additional phases were detected. The characteristic peaks of TiO2 appeared in the composite powders and the 7 wt% nano-TiO2 reinforced akermanite scaffold. All this meant that there were no phase change and reaction between akermanite and TiO2 after laser sintering.


image file: c4ra12095g-f4.tif
Fig. 4 XRD pattern of (a) the starting nano-TiO2 powders, (b) the starting akermanite powders, (c) the akermanite + 7 wt% nano-TiO2 composite powders, (d) the pure akermanite scaffold and (e) the 7 wt% nano-TiO2 reinforced akermanite scaffold.

3.2 Mechanical properties of the scaffolds

A representative Vickers indentation on the surface of strut is shown in Fig. 5a. The Vickers indentation diagonals and cracks at the tips of the indentation can be seen clearly. The cracks emanate from the four vertices along the indentation diagonals. The Vickers hardness and fracture toughness are determined by measuring the length of the diagonals and the cracks, respectively. The typical stress–strain curves of the 5 wt% nano-TiO2 reinforced porous akermanite scaffold are shown in Fig. 5b. The analysis of the stress and strain curves indicates that all scaffolds present a brittle behavior, which is characteristic for ceramic materials. The stiffness is determined by the slope of the initial linear section of the stress–strain curve. The average result of five compressive tests carried out under the same conditions for each type of scaffold is obtained.
image file: c4ra12095g-f5.tif
Fig. 5 (a) Representative Vickers indentation obtained on the surface of strut, (b) typical stress–strain curve of the 5 wt% nano-TiO2 reinforced porous akermanite scaffold.

Mechanical properties of the scaffolds in terms of fracture toughness, Vickers hardness, compressive strength and stiffness were investigated. The fracture toughness and Vickers hardness of the scaffolds as a function of nano-TiO2 are presented in Fig. 6a. It could be seen that the fracture toughness was improved considerably by the addition of nano-TiO2. A maximum fracture toughness of 2.32 MPa m1/2 was achieved for the scaffold with 7 wt% nano-TiO2 content, which is much higher than that of the pure akermanite scaffold (1.82 MPa m1/2). The Vickers hardness of the pure akermanite scaffold is 5.82 GPa, and it increases to 7.66 GPa when the content of nano-TiO2 is 7 wt%. The compressive strength and stiffness of scaffolds as a function of nano-TiO2 are shown in Fig. 6b. It can be observed that the compressive strength and stiffness increase from 3.53 MPa to 22.86 MPa and 31.09 MPa to 122.36 MPa with increasing nano-TiO2 from 0 to 7 wt%, respectively. The mechanical properties of the scaffolds could be significantly improved by addition of nano-TiO2.


image file: c4ra12095g-f6.tif
Fig. 6 (a) Effect of nano-TiO2 on the fracture toughness and Vickers hardness of scaffolds, (b) effect of nano-TiO2 on the compressive strength and stiffness of scaffolds.

The mechanical properties increased rapidly with increasing nano-TiO2 addition from 0 wt% to 5 wt%, while it increased slowly with further increase in nano-TiO2 (up to 7 wt%). The slow increase of mechanical properties was mainly due to the nano-TiO2 powders were not dispersed well in the akermanite matrix when nano-TiO2 increased from 5 wt% to 7 wt%. Furthermore, the presence of residual internal stresses resulting from the thermal expansion mismatch between akermanite and TiO2 could also decrease the mechanical properties. Previous studies44,45 showed that calcium phosphate bioceramics (e.g., HAP and β-TCP) have good biocompatibility with human body, but the low mechanical properties are the main limitation in load-bearing applications. Some reports have showed that the fracture toughness of HAP and β-TCP was 0.8–1.2 and 0.95–1.21 MPa m1/2, respectively.46,47 In this study, the fracture toughness of pure akermanite scaffold is 1.82 MPa m1/2, which is much higher than those of calcium phosphate bioceramics. The mechanical properties of the scaffolds are significantly increased with the content of nano-TiO2 increasing from 0 to 7 wt%. The highest fracture toughness (2.32 MPa m1/2) of the reinforced porous akermanite scaffolds is slightly higher than the lowest value of cortical bone (2–12 MPa m1/2),48 and the highest compressive strength (22.86 MPa) is much lower than cortical bone (130–180 MPa).49 Previous study showed that the fracture toughness of akermanite ceramics is 1.83 MPa m1/2 sintered at 1370 °C for 6 h,10 and the compressive strength is between 1.13 to 0.53 MPa with the porosity from 63.5 to 90.3%, respectively.50

The SEM/EDS analysis of starting akermanite powders and nano-TiO2 powders are shown in Fig. 7a and b, respectively. The starting akermanite powders have irregular shape and particle size of about 1–5 μm (Fig. 7a). The starting nano-TiO2 powders are spherical and slightly agglomerated with an average particle size of 20 nm (Fig. 7b). The corresponding EDS analysis shows that Ca, O, Mg and Si are present on the starting akermanite powders, while only Ti and O are present on nano-TiO2 powders. The SEM/EDS analysis of the polished and chemically etched surface of the pure akermanite and 5 wt% nano-TiO2 reinforced akermanite are shown in Fig. 7c and d, respectively. The average grain size was determined from SEM images with the linear intercept method. It can be seen that the micro size grains appearance with clear grain boundaries and the average grain size is about 5.3 μm (Fig. 7c). The EDS analysis of the pure akermanite after etching reveals the presence of Ca, O, Mg and Si, agreeing with the elemental composition of akermanite. Some nano-TiO2 is dispersed within micro-sized akermanite grains, others are found in the intergranular region (Fig. 7d). The average grain size of the reinforced scaffold is about 3.3 μm. It indicates that the nano-TiO2 can inhibit the grain growth by pinning effect of grain boundaries.51 The strength of ceramic will increase with decreasing of grain size according to the Hall–Petch relationship. The EDS analysis confirms the presence of nano-TiO2 in the akermanite grains as well as along the grain boundaries.


image file: c4ra12095g-f7.tif
Fig. 7 SEM/EDS of (a) the starting akermanite powders, (b) the starting nano-TiO2 powders, (c) the pure akermanite after etching and (d) the 5 wt% nano-TiO2 reinforced akermanite after etching.

The location of nano size grains in the micro size grains matrix can be divided into three types according to the classification of Niihara52 as illustrated in Fig. 8. In inter or intra type, the nano size grains are dispersed mainly at the grain boundaries of the matrix grains (Fig. 8a) or within the matrix (Fig. 8b). The inter/intra type with nano size grains are distributed both within the micro size grains and along the grain boundaries (Fig. 8c). According to this classification, the microstructure type of the nano-TiO2 reinforced akermanite is denoted as the inter/intra type. Sawaguchi et al.53 pointed out that the inter/intra type possess the highest toughness than the other two types.


image file: c4ra12095g-f8.tif
Fig. 8 Three types of nanocomposites: (a) inter type, that is the nano size grains are located at the grain boundaries of micro size grains. (b) Intra type, that is the nano size grains are distributed within the micro size grains. (c) Inter/intra type, that is the nano size grains are distributed at the grain boundaries as well as within the micro size grains.

The Vickers indentation crack propagation in the pure akermanite and nano-TiO2 reinforced akermanite is shown in Fig. 9. The crack in the pure akermanite propagates relatively straightly (Fig. 9a). However, the crack in the nano-TiO2 reinforced akermanite propagates flexurally, which inclines and distorts many times (Fig. 9b). The change of crack direction increases the length of crack path and reduces the stress intensity near crack tip.54 The crack is propagating mainly around the nano-TiO2 particles, occasionally end at the hard particle when the crack tip encounters nano-TiO2 particles (Fig. 9c and d). These processes can absorb additional amounts of fracture energy, which could apparently enhance the fracture toughness of nano-TiO2 reinforced akermanite. This investigation is also supported by the mechanical properties tests results in Fig. 6. The thermal expansion coefficient of akermanite and TiO2 are about 32.1 × 10−6 °C−1 (ref. 55) and 7 × 10−6 °C−1,56 respectively. Residual stress generated around the nano-TiO2 particles was due to the thermal expansion mismatch when the reinforced akermanite cooled from the high sintering temperature. The large residual stress gradient is located at the interface between relatively hard nano-TiO2 particles and soft akermanite matrix. When the crack propagating along the boundaries encounters the nano-TiO2 hard particles, it can be deflected from its original trajectory, where the crack would be impeded by the intragranular particles and deflected again.


image file: c4ra12095g-f9.tif
Fig. 9 Crack propagation in the (a) pure akermanite and (b–d) nano-TiO2 reinforced akermanite.

Meanwhile, the internal stress produced by the nano-TiO2 particles at the grain boundaries tends to alter the crack propagation path from intergranular fracture to transgranular fracture.57 The fracture surfaces of akermanites with different nano-TiO2 are shown in Fig. 10. The fracture mode of the pure akermanite is intergranular fracture (Fig. 10a). It changes from intergranular fracture to more and more transgranular fracture (Fig. 10b–d) with nano-TiO2 increasing. The fracture mode of akermanites reinforced with 5 wt% nano-TiO2 was mainly transgranular fracture (Fig. 10d). The transgranular fracture mode significantly helps to strengthen the grain boundaries and resist against crack propagation. It can be concluded that the addition of nano-TiO2 particles in akermanite matrix could increase the toughness and strength of porous akermanite scaffolds.


image file: c4ra12095g-f10.tif
Fig. 10 Fracture surfaces of akermanites with different nano-TiO2 (a) 0 wt%, (b) 1 wt%, (c) 3 wt%, (d) 5 wt%.

3.3 Bioactivity and biocompatibility of the scaffolds

It is known that TiO2 has been considered as a bioinert ceramic,58 so its content in porous scaffold should be as low as possible. As the mechanical properties of 5 wt% nano-TiO2 reinforced akermanite scaffold was close to that of 7 wt% nano-TiO2 reinforced akermanite scaffold, we chose the scaffold with 5 wt% nano-TiO2 to carry out the biological properties tests. The surface morphologies of the 5 wt% nano-TiO2 reinforced porous akermanite scaffold after immersion in SBF for 2 and 6 days were shown in Fig. 11a, b, d and e, respectively. It could be seen that the surface of the scaffold was partially covered by some particles after 2 days of immersion in SBF (Fig. 11a). The newly formed particles were worm-like (Fig. 11b). The EDS results revealed that there was obvious signal of P element for the mineralized surface of the scaffold. The deposited particles possessed a higher Ca/P ratio (Ca/P = 2.33) than that of HAP (Ca/P = 1.67) (Fig. 10c). The scaffold surface was fully covered by a thick precipitate layer after immersing in SBF for 6 days (Fig. 11d). The formed precipitate had a sponge-like structure (Fig. 11e). The Ca/P ratio of the layer was about 1.75, which was close to that of HAP (Fig. 11f). The XRD patterns of the scaffolds after soaking in SBF for 2 and 6 days were shown in Fig. 11g and h, respectively. The crystalline peak of apatite at 25.7°, 31.8°, 32.7° and 34° were evident in the pattern,59 which indicated that akermanite scaffolds induced the apatite formation when soaked in SBF.
image file: c4ra12095g-f11.tif
Fig. 11 SEM micrographs of the 5 wt% nano-TiO2 reinforced porous akermanite scaffolds soaked in SBF for 2 days (a and b) and 6 days (d and e); (c and f) the EDS of (b and e) in a selected region, XRD patterns of the scaffolds after soaking in SBF for 2 days (g) and 6 days (h).

The cross-sectional views of the akermanite scaffold with and without the 5 wt% nano-TiO2 after soaking in SBF for 2 and 6 days were shown in Fig. 12. It could be seen that the all the scaffolds surfaces were covered by the homogenous three-dimensional (3D) lamellar structure, which was the bone like apatite layer. The thickness of the apatite layer increased obviously with the increase of soaking time. No significant differences were observed in the cross-sectional of the akermanite scaffold with and without 5 wt% nano-TiO2 after soaking in SBF for the same time. The results indicated that the nano-TiO2 did not alter the apatite forming ability of akermanite scaffolds. Previous studies60,61 have shown that the bone like apatite plays an important role in inducing bone formation and enhancing cell adhesion and proliferation. The formation of bone like apatite suggested that the nano-TiO2 reinforced akermanite porous scaffold possessed good bioactivity.


image file: c4ra12095g-f12.tif
Fig. 12 Cross-sectional views observed by SEM of the pure akermanite scaffold (a and b) and the 5 wt% nano-TiO2 reinforced akermanite scaffold (c and d) soaked in SBF for 2 days (a and c) and 6 days (b and d).

The cell viability of MG-63 cells cultured on the akermanite scaffolds with and without the 5 wt% nano-TiO2 after 3, 7, and 10 days were shown in Fig. 13. Obvious cell proliferation was observed on all of the scaffolds, indicating that all the materials were biocompatible and could support cell proliferation. The viability of MG-63 cells on the scaffolds was statistically lower at the early days but increased dramatically with prolonging the culture time. MG-63 cells proliferated more favorably in the akermanite scaffolds with the 5 wt% nano-TiO2 than the scaffolds made of akermanite only, which indicated that the addition of 5 wt% nano-TiO2 could improve the biocompatible of akermanite scaffolds. Studies had shown that TiO2 was able to enhance osteoblast adhesion and could induce cell growth.62–64


image file: c4ra12095g-f13.tif
Fig. 13 MG-63 cells numbers based on an MTT assay within the porous akermanite scaffolds. Error bars represent a standard error of mean values where n = 5.

The light-microscopic morphology of original MG-63 cells viewed under light microscope was shown in Fig. 14a. The original MG-63 cells were fusiform or polygon shaped and uniform in size. The morphologies of MG-63 cultured on the pure akermanite scaffold and the 5 wt% nano-TiO2 reinforced porous akermanite scaffold for different days were shown in Fig. 14b–d and e–g, respectively. The MG-63 cells attached and spread well on the surface of the scaffold after culturing for 3 days (Fig. 14b and e). The number of MG-63 cells attached on the scaffold surface increased significantly with increasing of culture time (Fig. 14c–d and f–g). The cells were flattened and attached tightly on the scaffolds surfaces with their filopodium and lamellipodium, suggesting good cell viability on all akermanite scaffold. Previous studies50 showed that the Ca, Si and Mg ions from akermanite scaffolds dissolution could stimulate MG-63 cells proliferation and this stimulatory effect might be considered as one of the evaluation criteria for bioactivity of the scaffolds. Our results indicated that the nano-TiO2 reinforced akermanite scaffold possessed excellent biocompatibility to enhance cell attachment and growth.


image file: c4ra12095g-f14.tif
Fig. 14 Light-microscopic morphology of original MG-63 cells (a); SEM micrographs of the MG-63 cells (arrows) on the pure akermanite scaffold (b–d) and the 5 wt% nano-TiO2 reinforced akermanite scaffold (e–g) after cell culture for 3 days (b and e), 7 days (c and f) and 10 days (d and g).

4 Conclusions

The nano-TiO2 reinforced akermanite scaffolds with an open, uniform and interconnected pore structure were successfully fabricated by SLS at the laser power of 7.8 W, laser scan speed of 100 mm min−1 and spot size of 1.2 mm. The results showed that the Vickers hardness, fracture toughness, compressive strength and stiffness increased rapidly from 5.82 GPa to 7.57 GPa, 1.82 to 2.27 MPa m1/2, 3.53 to 21.81 MPa and 31.09 to 111.05 MPa as nano-TiO2 increased from 0 to 5 wt%, and then increased slowly to 7.66 GPa, 2.32 MPa m1/2, 22.86 MPa and 122.36 MPa as nano-TiO2 increased to 7 wt%, respectively. The mechanisms of remarkable improvement in the mechanical properties were observed to be grain size refinement, crack deflection and fracture mode transition. The akermanite scaffolds reinforced with 5 wt% nano-TiO2 possessed excellent ability to form bone like apatite in SBF and support cell attachment and growth. It is concluded that the porous akermanite scaffold with 5 wt% of nano-TiO2 may be a good candidate for bone tissue engineering applications.

Acknowledgements

This work was supported by the following funds: (1) The Natural Science Foundation of China (51222506, 81372366, 81472058); (2) Overseas, Hong Kong & Macao Scholars Collaborated Researching Fund of National Natural Science Foundation of China (81428018); (3) Hunan Provincial Natural Science Foundation of China (14JJ1006); (4) Project supported by the Fok Ying-Tong Education Foundation, China (131050); (5) Shenzhen Strategic Emerging Industrial Development Funds (JCYJ20130401160614372); (6) The Open-End Fund for the Valuable and Precision Instruments of Central South University; (7) The faculty research grant of Central South University (2013JSJJ011, 2013JSJJ046); (8) State Key Laboratory of New Ceramic and Fine Processing Tsinghua University (KF201413); (9) The Fundamental Research Funds for the Central Universities of Central South University.

Notes and references

  1. H. Gu, F. Guo, X. Zhou, L. Gong, Y. Zhang, W. Zhai, L. Chen, L. Cen, S. Yin and L. Cui, Biomaterials, 2011, 32, 7023 CrossRef CAS PubMed.
  2. M. Yeo, C. G. Simon and G. H. Kim, J. Mater. Chem., 2012, 22, 21636 RSC.
  3. W. Zhai, H. Lu, L. Chen, X. Lin, Y. Huang, K. Dai, K. Naoki, G. Chen and J. Chang, Acta. Biomater., 2012, 8, 341 CrossRef CAS PubMed.
  4. E. Abed and R. Moreau, Cell Proliferation, 2007, 40, 849 CrossRef CAS PubMed.
  5. C. Wu, Y. Ramaswamy and H. Zreiqat, Acta. Biomater., 2010, 6, 2237 CrossRef CAS PubMed.
  6. Z. Han, C. Gao, P. Feng, Y. Shen, C. Shuai and J. Wang, RSC Adv., 2014, 4, 36868 RSC.
  7. C. Wu and J. Chang, J. Biomed. Mater. Res., Part B, 2007, 83, 153 CrossRef PubMed.
  8. X. Hou, G. Yin, X. Chen, X. Liao, Y. Yao and Z. Huang, Appl. Surf. Sci., 2011, 257, 3417 CrossRef CAS PubMed.
  9. W. Zhai, H. Lu, C. Wu, L. Chen, X. Lin, K. Naoki, G. Chen and J. Chang, Acta. Biomater., 2013, 9, 8004 CrossRef CAS PubMed.
  10. C. Wu and J. Chang, J. Biomater. Appl., 2006, 21, 119 CrossRef CAS PubMed.
  11. Q. Fu, E. Saiz, M. N. Rahaman and A. P. Tomsia, Mater. Sci. Eng., C, 2011, 31, 1245 CrossRef CAS PubMed.
  12. H. Yang, S. Wu, J. Hu, Z. Wang, R. Wang and H. He, Mater. Des., 2011, 32, 1590 CrossRef CAS PubMed.
  13. E. Fidancevska, G. Ruseska, J. Bossert, Y. Lin and A. R. Boccaccini, Mater. Chem. Phys., 2007, 103, 95 CrossRef CAS PubMed.
  14. S. Kim, Y. Kong, I. Lee and H. Kim, J. Mater. Sci.: Mater. Med., 2002, 12, 307 CrossRef.
  15. D. Lahiri, S. Ghosh and A. Agarwal, Mater. Sci. Eng., C, 2012, 32, 1727 CrossRef CAS PubMed.
  16. K. Balani, R. Anderson, T. Laha, M. Andara, J. Tercero, E. Crumpler, J. Tercero, E. Crumpler and A. Agarwal, Biomaterials, 2007, 28, 618 CrossRef CAS PubMed.
  17. Y. Chen, Y. Q. Zhang, T. H. Zhang, C. H. Gan, C. Y. Zheng and G. Yu, Carbon, 2006, 44, 37 CrossRef CAS PubMed.
  18. Z. Huan, S. Leeflang, J. Zhou, W. Zhai, J. Chang and J. Duszczyk, J. Biomed. Mater. Res., Part B, 2012, 100, 437 CrossRef PubMed.
  19. M. Uchida, H. M. Kim, T. Kokubo, S. Fujibayashi and T. Nakamura, J. Biomed. Mater. Res., Part A, 2003, 64, 583 Search PubMed.
  20. K. Lin, J. Chang, Z. Liu, Y. Zeng and R. Shen, J. Eur. Ceram. Soc., 2009, 29, 2937 CrossRef CAS PubMed.
  21. H. Tiainen, D. Wiedmer and H. J. Haugen, J. Eur. Ceram. Soc., 2013, 33, 15 CrossRef CAS PubMed.
  22. R. Sabetrasekh, H. Tiainen, S. P. Lyngstadaas, J. Reseland and H. Haugen, J. Biomater. Appl., 2011, 25, 559 CrossRef CAS PubMed.
  23. K. A. Khalil, S. W. Kim, K. W. Kim, N. Dharmaraj and H. Y. Kim, Int. J. Appl. Ceram. Technol., 2011, 8, 523 CrossRef CAS PubMed.
  24. C. Kailasanathan and N. Selvakumar, Ceram. Int., 2012, 38, 3569 CrossRef CAS PubMed.
  25. Z. Seeley, A. Bandyopadhyay and S. Bose, J. Biomed. Mater. Res., Part A, 2007, 82, 113 CrossRef PubMed.
  26. S. Ahn, Y. Kim, H. Lee and G. Kim, J. Mater. Chem., 2012, 22, 15901 RSC.
  27. C. G. Jeong and S. J. Hollister, Tissue Eng., Part A, 2010, 16, 3759 CrossRef CAS PubMed.
  28. C. Wu, L. Chen, J. Chang, L. Wei, D. Chen and Y. Zhang, RSC Adv., 2013, 3, 17843 RSC.
  29. W. Y. Yeong, N. Sudarmadji, H. Y. Yu, C. K. Chua, K. F. Leong, S. S. Venkatraman, Y. C. Boey and L. P. Tan, Acta. Biomater., 2010, 6, 2028 CrossRef CAS PubMed.
  30. M. G. Li, X. Y. Tian and X. B. Chen, Biofabrication, 2009, 1, 032001 CrossRef CAS PubMed.
  31. C. Shuai, P. Feng, C. Cao and S. Peng, Biotechnol. Bioprocess., 2013, 18, 520 CrossRef CAS.
  32. Y. Liu, F. Zhu, X. Dong and W. Peng, J. Zhejiang Univ., Sci., B, 2011, 12, 769 CrossRef CAS PubMed.
  33. A. Krishnan and L. R. Xu, J. Nanomater., 2012, 2012, 1 CrossRef PubMed.
  34. J. L. Drummond, J. Dent. Res., 2008, 87, 710 CrossRef CAS PubMed.
  35. C. Domingo, R. W. Arcís, A. López-Macipe, R. Osorio, R. Rodríguez-Clemente, J. Murtra, M. A. Fanovich and M. Toledano, J. Biomed. Mater. Res., 2001, 56, 297 CrossRef CAS.
  36. C. Shuai, C. Gao, Y. Nie, H. Hu, Y. Zhou and S. Peng, Nanotechnology, 2011, 22, 1 CrossRef PubMed.
  37. L. Siggelkow, U. Burkhardt, G. Kreiner, M. Palm and F. Stein, Mater. Sci. Eng., A, 2008, 497, 174 CrossRef PubMed.
  38. A. G. Evans and E. A. Charles, J. Am. Ceram. Soc., 1976, 59, 371 CrossRef CAS PubMed.
  39. T. Kokubo, S. Ito, Z. T. Huang, T. Hayashi, S. Sakka, T. Kitsugi and T. Yamamuro, J. Biomed. Mater. Res., 1990, 24, 331 CrossRef CAS PubMed.
  40. C. Shuai, P. Feng, L. Zhang, C. Gao, H. Hu, S. Peng and A. Min, Sci. Technol. Adv. Mater., 2013, 14, 055002 CrossRef.
  41. H. E. Gotz, M. Muller, A. Emmel, U. Holzwarth, R. G. Erben and R. Stangl, Biomaterials, 2004, 25, 4057 CrossRef CAS PubMed.
  42. S. J. Hollister, C. Y. Lin, E. Saito, R. D. Schek, J. M. Taboas, J. M. Williams, B. Partee, C. L. Flanagan, A. Diggs, E. N. Wilke, G. H. Van Lenthe, R. Müller, T. Wirtz, S. Das, S. E. Feinberg and P. H. Krebsbach, Orthod. Craniofac. Res., 2005, 5, 162 CrossRef PubMed.
  43. T. D. Roy, J. L. Simon, J. L. Ricci, E. D. Rekow, V. P. Thompson and J. R. Parsons, J. Biomed. Mater. Res., Part A, 2003, 66, 283 CrossRef PubMed.
  44. S. Eshraghi and S. Das, Acta. Biomater., 2012, 8, 3138 CrossRef CAS PubMed.
  45. Y. Zhang, D. Kong and X. Feng, Ceram. Int., 2012, 38, 2991 CrossRef CAS PubMed.
  46. V. P. Orlovskii, V. S. Komlev and S. M. Barinov, Inorg. Mater., 2002, 38, 1159 Search PubMed.
  47. D. Liu, J. Zhuang, C. Shuai and S. Peng, Biofabrication, 2013, 5, 025005 CrossRef PubMed.
  48. S. Ramesh, C. Y. Tan, I. Sopyan, M. Hamdi and W. D. Teng, Sci. Technol. Adv. Mater., 2007, 8, 124 CrossRef CAS PubMed.
  49. D. Espalin, K. Arcaute, D. Rodriguez, F. Medina, M. Posner and R. Wicker, Rapid Prototyping Journal, 2010, 16, 164 CrossRef.
  50. C. Wu, J. Chang, W. Zhai, S. Ni and J. Wang, J. Biomed. Mater. Res., Part B, 2006, 78, 47 CrossRef PubMed.
  51. Y. Yang, P. Lan, M. Wang, T. Wei, R. Tan and W. Song, Nanoscale Res. Lett., 2012, 7, 1 CrossRef PubMed.
  52. K. Niihara, J. Ceram. Soc. Jpn., 1991, 99, 974 CrossRef CAS.
  53. A. Sawaguchi, K. Toda and K. Niihara, J. Am. Ceram. Soc., 1991, 74, 1142 CrossRef CAS PubMed.
  54. Y. Yang, Y. Wang, W. Tian, Z. Wang, Y. Zhao, L. Wang and H. Bian, Mater. Sci. Eng., A, 2009, 508, 161 CrossRef PubMed.
  55. D. Yi, C. Wu, X. Ma, H. Ji, X. Zheng and J. Chang, Biomed. Mater., 2012, 7, 065004 CrossRef PubMed.
  56. X. Miao, D. Sun and P. W. Hoo, Ceram. Int., 2009, 35, 281 CrossRef CAS PubMed.
  57. K. A. Khalil and H. Y. Kim, Int. J. Appl. Ceram. Technol., 2007, 4, 30 CrossRef CAS PubMed.
  58. F. Z. Mezahi, A. Harabi, S. Zouai, S. Achour and D. Bernache-Assollant, Mater. Sci. Forum, 2005, 492, 241 CrossRef.
  59. X. Qi, H. Li, B. Qiao, W. Li, X. Hao, J. Wu, B. Su and D. Jiang, Int. J. Nanomed., 2013, 8, 4441 CrossRef PubMed.
  60. C. T. Wu, J. Chang, J. Y. Wang, S. Y. Ni and W. Y. Zhai, Biomaterials, 2005, 26, 2925 CrossRef CAS PubMed.
  61. G. Lehmann, H. Cacciotti, P. Palmero, L. Montanaro, A. Bianco, L. Campagnolo and A. Camaioni, Biomed. Mater., 2012, 7, 1 Search PubMed.
  62. W. Xu, W. Y. Hu, M. H. Li and C. Wen, Mater. Lett., 2006, 60, 1575 CrossRef CAS PubMed.
  63. S. Nath, R. Tripathi and B. Basu, Mater. Sci. Eng., C, 2009, 29, 97 CrossRef CAS PubMed.
  64. S. Pushpakanth, B. Srinivasan, B. Sreedhar and T. P. Sastry, Mater. Chem. Phys., 2008, 107, 492 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.