Effect of Sm2O3 substitution on mechanical and

3 downloads 0 Views 7MB Size Report
May 11, 2018 - J Aust Ceram Soc (2018) 54:621-630. DOI 10.1007/s41779-018-0190-7. Effect of Sm2O3 substitution on mechanical and biological properties ...
Effect of Sm2O3 substitution on mechanical and biological properties of 45S5 bioactive glass Md Ershad, Vikas Kumar Vyas, Sunil Prasad, Akher Ali & Ram Pyare

Journal of the Australian Ceramic Society ISSN 2510-1560 Volume 54 Number 4 J Aust Ceram Soc (2018) 54:621-630 DOI 10.1007/s41779-018-0190-7

1 23

Your article is protected by copyright and all rights are held exclusively by Australian Ceramic Society. This e-offprint is for personal use only and shall not be self-archived in electronic repositories. If you wish to self-archive your article, please use the accepted manuscript version for posting on your own website. You may further deposit the accepted manuscript version in any repository, provided it is only made publicly available 12 months after official publication or later and provided acknowledgement is given to the original source of publication and a link is inserted to the published article on Springer's website. The link must be accompanied by the following text: "The final publication is available at link.springer.com”.

1 23

Author's personal copy Journal of the Australian Ceramic Society (2018) 54:621–630 https://doi.org/10.1007/s41779-018-0190-7

RESEARCH

Effect of Sm2O3 substitution on mechanical and biological properties of 45S5 bioactive glass Md Ershad 1

&

Vikas Kumar Vyas 1 & Sunil Prasad 1 & Akher Ali 1 & Ram Pyare 1

Received: 13 September 2017 / Revised: 9 March 2018 / Accepted: 10 April 2018 / Published online: 11 May 2018 # Australian Ceramic Society 2018

Abstract Synthesis and characterization of bioactive glasses having a general composition of (45 − X) SiO2, 24.5 Na2O, 24.5 CaO, and 6 P2O5 (where X = 0–4.0 wt% of Sm2O3) were studied. These glasses were melted in the alumina crucible at a temperature of 1400 ± 5 °C with air as a furnace atmosphere. Differential thermal analysis (DTA) was carried out at a heating rate of 10 °C per min under argon gas atmosphere using alumina as reference material in the temperature range of 100 to 1000 °C. From the DTA curve, nucleation and crystal growth temperature was determined and found to decrease with increasing content of Sm2O3. The glass samples were immersed in simulated body fluid (SBF) for different time periods, and bioactivities were measured using FTIR and pH. It was observed that the bioactivity of glass increased with increasing concentration of Sm2O3. Microstructures were evaluated by a scanning electron microscope (SEM). Density and other mechanical properties such as flexural, compressive, and microhardness were measured and were found to increase with increasing concentration of Sm2O3. Keywords Bioactive glass . DTA . XRD . FTIR . Scanning electron microscope (SEM)

Introduction A bioactive material is defined as a material that stimulates a biological response from the body, mainly bonding to soft as well as hard tissues. The name Bbio-ceramic^ is a common term used to cover glass ceramics, and they are used as implant materials like bone and teeth. The main discovery of the bioglass composition was as follows (mol%): 46.1 SiO2, 24.4 Na2O, 26.9 cows, and 2.6 P2O5, and the weight percent composition of 45% SiO2, in weight % with network modifiers of 24.5% Na2O and 24.5% CaO and 6% P2O5, was determined to make for the glass composition [1]. This first-generation biomaterials move toward the replacement of tissues that was irreversibly altered when a special composition of soda– lime–phosphate–silicate glass was made by implanting in the femurs of rats in 1969 by Hench and colleagues at the University of Florida [2–4]. The key shortcoming of 45S5 bioactive glass is connected tissue to its slow degradation rate.

* Md Ershad [email protected] 1

Department of Ceramic Engineering, Indian Institute of Technology (Banaras Hindu University), Varanasi 221005, India

According to some past results, the mechanical properties of 45S5 bioactive glass were not wholly adequate for significant load-bearing application [5]. Bioactive glasses attach to the living bone through an apatite layer formed on their surfaces after being dipped into a simulated body fluid (SBF) [6]. The bioactive material should possess excellent biochemical behavior and biomechanical strength. The 45S5® bioactive glass has a very good capability to bond with both soft and hard tissues [7]. Hench decided to make a glass of the SiO2–Na2O–CaO–P2O5 system containing high calcium contents with a composition close to a ternary eutectic in the Na2O–CaO–SiO2 diagram [8]. The main discovery was a glass of the mol% composition 46.1 SiO2– 24.4 Na2O–26.9 CaO–2.6 P2O5, known as the 45S5 bioglass, which formed a strong bond with bone which was not removed easily without breaking the bone [1]. Substitution of bioactive glasses with different transition and rare-earth metals such as cerium, gadolinium, lanthanum, zinc, manganese, iron, magnesium, and silver to change their biological and bioactive responses has been studied by many research groups [9–15]. The Sm3+ ions play the important role as a network modifier in the Sm2O3–SiO2–Na2O–CaO–P2O5 glass system, as Sm3+ possesses strong fluorescence intensity, high bending energy levels (135.6 BE/eV), and high quantum efficiency which further enhances optical as well as

Author's personal copy 622

J Aust Ceram Soc (2018) 54:621–630

mechanical properties of a glass system [16–19]. A biomaterial is a synthetic material to be used in the intimate contact with living tissue. In this work, we study the effect of Sm2O3 doping on the thermal, optical, mechanical, and in vitro properties of the bioactive glass. Sm3+ ions have been used in dental ceramics to mimic the fluorescence of natural teeth [20]. Samarium is also known to possess bacteriostatic properties and has low toxicity [21]. The 45S5 bioactive glass is a very successful biomaterial for clinical applications, and many researchers have studied an incorporation of some ions such as Co, Ni, Li, Ti, K, Zr, Mg, and Sr in the base bioactive glass because of their unique outcome on osteoblastic cell proliferation of different ions in the base bioactive glasses [22–27]. The present work is concerned with the preparation and characterization of samarium-containing bioactive glass (BG). These bioactive glasses were immersed in SBF to validate the formation of a bone-like apatite layer on their surfaces by using FTIR. Surface morphology was studied using scanning electron microscopy (SEM), and mechanical properties were studied in the present investigation.

Experimental Materials and methods The wt% compositions of Sm2O3 (99.9% pure) containing 45S5 bioglass samples are given in Table 1. Fine-grained quartz (99.9%) were used as a source of SiO2. Analytical reagent grades, CaCO3, Na2CO3, and (NH4)H2PO4, were used as a source of CaO, Na2O, and P2O5, respectively. The required amounts of analytical reagent grade Sm2O3 were added one-by-one in the batch given in Table 1, for the partial substitution of SiO2. The raw materials for different samples were properly weighed. Then, the mixing of different batches was done for 30 min in a ball mill of 480 rpm. After that, they were melted in a 100-ml alumina crucible (99.9%) and melted in a Globar furnace at 1400 ± 10 °C in the air as furnace atmosphere (Fig. 1). The temperature of the furnace was controlled within ± 5 °C by an automatic temperature indicator Table 1 Samples

Bg Sm1 Sm2 Sm3 Sm4

wt% composition of the bioglass samples wt% SiO2

Na2O

CaO

P2O5

Sm2O3

45.0 44.0 43.0 42.0 41.0

24.5 24.5 24.5 24.5 24.5

24.5 24.5 24.5 24.5 24.5

6.0 6.0 6.0 6.0 6.0

0.0 1.0 2.0 3.0 4.0

Fig. 1 Flowchart of the bioglass samples prepared by the glass melting route

controller. The temperature of a furnace was held for 3 h at 1400 °C. After that, melted samples were poured onto a preheated aluminum sheet and directly transferred to a regulated muffle furnace at 520 °C for annealing. After 2 h, the annealing furnace was switched off and glass was taken at room temperature. Some part of the glass was powdered using an agate pestle and mortar for DTA, X-ray diffraction (XRD), FTIR, and SEM analysis, and the rest was cut into square shapes for determining mechanical properties.

Preparation of SBF The prepared glass samples were immersed in SBF solution at 37.4 °C for different time periods varying from 1 to 21 days. The solution of SBF was prepared according to the procedure described by Kokubo and Takadama [28]. Table 2 shows the ion concentration of SBF solution. The pH of the SBF solution was measured using a digital pH meter after immersion of samples for different time periods.

Powder XRD The bioglass samples were ground to 50 μm, and the fine powders were subjected to XRD analysis using a Rigaku Miniflex II diffractometer which adopted Cu-Kα radiation (λ = 1.5405 Å) with a tube voltage of 40 kV and current of 35 mA in a 2θ range between 20° and 80° onward. The step size and scan speed were set at 0.02° and 2°/min, respectively. The JCPDS-International Centre for Diffraction Data Cards was used as a reference.

Structural and in vitro study of bioactive glass The structures of bioglass samples were measured at room temperature in the frequency range of 4000–400 cm−1 using

Author's personal copy J Aust Ceram Soc (2018) 54:621–630

623

Table 2 Ion concentration (mM/ l) of the simulated body fluid and human blood plasma

Ion Na+

Mg2+

HCO3−

Cl−

HPO42−

SO42−

142

5.0

2.5

1.5

4.2

148

1.0

0.5

Human blood plasma

142

5.0

2.5

1.5

27.0

103

1.0

0.5

pH and physicomechanical properties The bioglass powder (0.2 g) was soaked in 20 ml of SBF solution at 37 °C for a different time period, and the pH was measured by using the microprocessor-based pH EC meter (model 1611; Esico, USA). The density of the bioglass samples was determined by the Archimedes principle with water as the immersion liquid. The measurements were carried out at 25 °C. All the weight measurements were made using a digital balance (model BP221S; Sartorius, USA) with an accuracy of ± 0.0001 g. Density (ρ) of the sample was obtained by employing Eq. (1) as given below Wa ρ W a −W b b

Ca2+

Simulated body fluid

a Fourier transform infrared spectrometer (Bruker Tensor 27 FTIR, USA). In order to appraise the formation of (calcium phosphate) apatite layer on the surface of the samples after dipping in SBF solution, the glass samples (0.2 g) were immersed in 20 ml of SBF solution in a small plastic container at 37 °C. The pH of SBF solution was 7.40. These samples were put in an incubator in the static condition for a period of 1, 4, 8, 13, and 21 days. After soaking, the samples were filtered, rinsed with doubly distilled water, and dried in an oven at 100 ± 2 °C for 2.3 h before analysis by FTIR.

ρ¼

K+

ð1Þ

where Wa is the weight of the sample in air, Wb is the weight of the sample in buoyant liquid, and ρb is the density of the buoyant liquid. The mechanical strength of bioactive glass samples was measured by the universal testing machine (UTM). Ultrasonic measurement gauge determined Young’s modulus, shear modulus, and bulk modulus of glass samples. The ultrasonic wave velocities were recorded as longitudinal (VL) and transverse (VT) waves. The velocities of sound wave that passed in the polished bioglass samples were measured using ultrasonic pulse-echo techniques (EPOCH 600; Olympus, USA). The test was done using two transducers, one was V112 for longitudinal wave (10 MHz) and another was V156 for the transverse wave (5 MHz). Burnt honey was used as a bonding material between samples and transducers. Elastic properties such as Poisson’s ratio (σ), Young’s modulus (E), shear modulus (S), and bulk modulus (K) were calculated by the ASTM method.

Weight loss measurement The solubility of bioactive glasses was found by measuring weight loss in SBF at 37.6 °C kept inside the incubator. With the help of 400-grit Amery papers, the samples were polished. And so, they were rinsed in acetone for a moment and were put in little containers which contain SBF. After a certain interval of time up to 20 days, the samples were taken out, and with the help of tissue, excess moisture was removed. Then, the samples were dried and weighted. The percent changes in weight loss were directly correlated to glass corrosion or solubility in SBF. The weight loss was measured by Eq. (2) Weight loss ð%Þ ¼

W i −W f  100 Wi

ð2Þ

where Wi is the initial weight of the specimen and Wf is the weight on different days incubated in SBF solution.

Results and discussion Differential thermal analysis The thermal behavior curves of the bioactive glasses are shown in Fig. 2, and the results dictated that the incorporation of Sm3+ ions in the 45S5 bioactive glass reduced both the nucleation temperature from 601 to 598 °C and the crystallization temperature from 765 to 632 °C. Further, it was observed that the higher amount of modifiers in the composition has facilitated in lowering Tg point and the viscosity of the glass melt [29]. Substitution of Sm2O3 for SiO2 content in the composition caused the shift of exothermic peaks to lower temperatures as compared to that of the base glass. Therefore, lower energy is required to promote crystallization in the glass. This may be attributed to the presence of larger cations in the system which increased interference in glass network. The modifiers occupied the interstitial position in the glass structure and thereby weakened oxygen bond strength [30]. Arepalli et al. also discuss the process of thermal treatment; the high concentration of rare-earth ions in the glassy system has allowed early nucleation by the nucleating agents like P2O5 in the glass composition [29].

Author's personal copy 624

J Aust Ceram Soc (2018) 54:621–630

Fig. 2 DTA curve of the bioactive glass samples (Bg, Sm1, Sm2, Sm3, and Sm4)

XRD analysis and FTIR assessment

pH behavior in SBF

X-ray diffraction pattern of glass samples was prepared in Fig. 3. It was observed that no peaks were found which indicate that the bioglass sample was amorphous in nature. The Fourier transform infrared transmittance spectra of the bioactive glass samples were recorded in the frequency range of 400–4000 cm−1 and are shown in Fig. 4. The vibrational bands have been marked by the vertical dotted lines which were prepared similarly to transmission mode spectra, as given by earlier workers [31]. The parent bioactive glass revealed the sharp bands successively at about 497, 1397, 2175, and 3790 cm−1, indicating various functional groups. The transmittance spectral bands of bioactive glasses confirmed the main characteristic of the SiO4 tetrahedral silicate network, and it is attributed to the presence of SiO2 as a significant constituent [32]. The resultant FTIR spectra at just about 450 cm−1 are associated with a Si–O–Si symmetric bending mode, and the band at 775 cm−1 corresponded to Si–O–Si symmetric stretching of non-bridging oxygen atoms between SiO4 tetrahedra. The major broadband at about 1397 cm−1 was attributed to Si–O–Si asymmetric stretching. The minor sharp peak at 1630 cm−1 is associated due to the stretching mode of C–O vibration of CO32− groups. The infrared frequencies and related functional groups were also reported by El Batal and El Kheshen [22] in their bioglass ceramic systems. Hydroxyl (OH−) group peaks were found at 3506 cm−1 as shown in Fig. 4.

The variation of pH of bioactive glass samples after immersing in SBF solution up to 21 days is presented in Fig. 5. It was observed that for all bioactive glass samples, the pH increases

Fig. 3 XRD patterns of the bioactive glass samples before soaking them into the SBF solution

Author's personal copy J Aust Ceram Soc (2018) 54:621–630

625

Fig. 4 FTIR transmittance spectra of the bioactive glass samples (Bg, Sm1, Sm2, Sm3, and Sm4)

within 1–4 days as compared to the initial pH of the SBF solution at 7.4 under normal temperature and pressure condition. The increase in pH values is due to the fast release of cations through exchange with H+ or H3O+ ions in the SBF solution. The H+ ions were replaced by cations which caused an increase in hydroxyl concentration of the solution [33]. This leads to attack on the silica glass network, which results in silanol formation, leading to a decrease in pH after 3 days as

Fig. 5 Variation of pH of the bioactive glass samples after immersing in SBF up to 21 days

indicated in Fig. 5. When bioactive glass samples were immersed in SBF solution up to 13 days [34], the maximum pH value is 9.93 of base bioactive glass. This is due to the attack in the silica glass network, which results in silanol formation, leading to a decrease in pH as indicated in Fig. 5. It was also seen that the decrease in the pH of the SBF solution after 13 days was due to the breaking of glass network. The addition of Sm2O3 up to 4 wt% increases pH of the SBF solution containing immersed samples which obtain maxima after around 13 days, and then it decreased with time referring to the base glass sample. This dictates that the addition of Sm2O3 up to 3.0 wt% in the glass samples increases its bioactivity but, after 3 wt% of Sm 2 O 3 , retards the bioactivity of the glass samples. Morphological properties of bioactive glasses also indicate that soaking in SBF solution leads to the formation of hydroxyapatite layer on the surface of the samples [33, 34]. The maxima of pH values were recorded between 1 and 8 days at pH 9.93, 9.88, 9.79, 9.61, and 9.37 for the samples Bg–Sm4, respectively, at 37 °C under physiological condition, which is due to the fast dissolution rate. So, an increase in the pH value of SBF solution also favors the hydroxycarbonate apatite (HCA) formation. When the bone was formed, the cross-linking of the collagen chains and the subsequent precipitation of hydroxycarbonate apatite are pH dependent and require a high pH at the bone formation site [35].

Author's personal copy 626

J Aust Ceram Soc (2018) 54:621–630

Fig. 6 SEM of bioactive glasses Bg, Sm1, Sm2, Sm3, and Sm4 before immersion in SBF, respectively

Surface morphology of bioactive glass samples by SEM The surface morphology of bioactive glasses was analyzed before immersion in SBF solution by SEM and is shown in Fig. 6. The samples were gold coated before SEM analysis. A scanning electron microscope (Inspect

50 FEI) was used for determining the surface microstructure of bioactive glass samples. When the bioglass sample was immersed in the SBF solution for different time periods, the HCA layer was found on the surface of bioactive glass samples. The SEM of the bioglass surface pre-immersion reveals the form of many rod-shaped structures and irregular grains after the addition of

Fig. 7 SEM of bioactive glasses Bg, Sm1, Sm2, Sm3, and Sm4 after 21 days of immersion in SBF, respectively

Author's personal copy J Aust Ceram Soc (2018) 54:621–630

627

Fig. 8 Weight loss measurement of the bioactive glass in SBF solution up to 20 days

Sm2O3 as shown in Fig. 6, which are quite similar to the results observed by Tripathi et al. [33] and Hanan et al. [34]. Figure 7 shows the SEM micrographs of bioactive glass samples after soaking in SBF solution for 21 days. It is clear from the image that bioactive glass samples which were soaked in SBF solution for 21 days were covered with the irregular shape of HCA particles which have grown into agglomerates. With SBF treatment, Fig. 9 Density of bioglass

HCA clusters change in a finer structure after 21 days of soaking due to partial dissolution and re-precipitation phenomena in solution. The bioglass releases Ca2+ and Na+ ions from its surface via an exchange with the H3O+ ion in the SBF to form Si–OH or Sm–OH groups on their surfaces [34]. This happens due to solution refreshing and signifying the formation of a continuous layer of HCA [35]. After comparing these morphologies, it was

Author's personal copy 628

J Aust Ceram Soc (2018) 54:621–630

Fig. 10 Poisson’s ratio

Fig. 12 Shear modulus (GPa)

concluded that the micrographs show the formation of HCA on the surface of bioactive glass samples after immersion in SBF solution for 21 days. EDX analysis also confirmed the formation of HCA layer after immersing the samples into SBF solution after 21 days as shown in Fig. 7. Arepalli et al. also investigated the substitution of barium up to 1.6 mol% in the base bioactive glasses. In this work, the SEM image showed the formation of HCA layer on the surface of the bariumcontaining bioactive glasses after putting in the SBF [29]. In a single investigation, samarium-substituted glass reinforced hydroxyapatite with enhanced osteoblastic performance and antibacterial properties for bone tissue regeneration and also showed the Sm-substituted bioglasses exhibited improved osteoblastic cell response. The author found that Sm-doped 45S5 bioactive glass enhances the bioactive properties by the in vitro studies.

Weight loss measurement Figure 8 shows the wt% losses of bioactive glasses, the unsubstituted 45S5 glass showing less weight loss after soaking into SBF solutions. In all cases up to 20 days, the weight loss is increasing proportionally up to 1.54% with time and, after that, weight loss is nearly constant for each sample. From the pH graph, the inference is already taken out when the pH is increasing due to the release of Ca2+ and Na+ ions into SBF solution that means the weight loss in the Sm4 glass is more which is suitably mentioned by the weight loss graph. It was established that up to 4 days, the weight loss in BG–Sm4 glasses is about similar. That means the glasses with higher content of samarium possess the higher rate of dissolution, so the % weight loss is more.

Fig. 11 Young’s modulus (GPa)

Fig. 13 Bulk modulus (GPa)

Author's personal copy J Aust Ceram Soc (2018) 54:621–630

629

Mechanical testing

Conclusion

Density

In the present work, a comparative study was made on bioactive, physicochemical, and mechanical properties of multicomponent SiO2–Na2O–CaO–P2O5–Sm2O3 bioactive glasses with varying concentrations of SiO2. The following conclusions were drawn from these investigations:

Figure 9 shows the density of the glass samples as a function of Sm3+ within error bars. It was found that with an increase in wt% substitution of SiO2 by Sm2O3 in a base glass, the density of glass samples increases from 2.69 to 2.71 g/cm3, respectively, and this is due to the small amount of replacement of SiO2 with Sm2O3 and is attributed to the replacement of a light element (density of SiO2 − 2.65) with a heavier one (Sm2O3 − 8.35). This is attributed to the reason that the samarium ions have occupied interstitial sites within the glass network [35, 36]. Therefore, it increased the densities and resulted in creating new bonds with the incorporation of Sm3+ ions in the bioactive glasses. It caused the reinforcement of glass structure and resulted in improvement in the compression of the glass samples.

&

In the present investigation, it was found that Young’s shear and bulk moduli increase with increasing concentration of Sm2O3 in non-charge balance (NCB) glasses whereas Poisson’s ratio decreases with increasing concentration of the same oxides. After the addition of Sm2O3, the nucleation and crystallization temperature reduces. In substitution to Sm3+ ions, it was found that physical properties, as well as biological properties, increased with increasing concentration of Sm2O3. The XRD pattern of the modified 45S5 glass shows the amorphous phase behavior in the presence of Sm3+ ions. SEM and EDX analyses showed that the HCA layer is present on the surface of samples after immersion in SBF solutions. FTIR spectra showed the structural behavior of base and Sm2O3-substituted bioglass. Higher content of samarium possesses the higher rate of dissolution, so the % weight loss is more.

& & &

Elastic modulus Figures 10, 11, 12, and 13 represent the experimental values of elastic moduli, Young’s modulus (E), shear modulus (S), and bulk modulus (K) of the bioactive glass samples which increase with increasing concentration of Sm2O3. The elastic moduli of the bioactive glass samples showed similar trends regarding improvement in their mechanical properties with the variation in the ultrasonic velocities. The elastic modulus of alkali silicate glasses was normally found to increase with increasing concentration of the modifier above a critical limit as a result of an increase in cohesion [36]. Thus, a greater bulk modulus of the glass samples was partially attributed to the addition of more amounts of modifiers like Na2O and CaO in the silicate glass samples. The authors further mentioned that the effect of P2O5 on the elasticity of their glasses was not clear, although phosphorus was causing a more polymerized silicate network. Hench and Thompson [9] have pointed out that Young’s modulus of cortical bone is about 7–30 GPa, which is far below than most ceramic-based implants. Hence, for biomechanical compatibility, the SiO2–CaO– P2O5–Na2O–CeO2–Sm2O3 bioactive glass samples were found to be better. It was also known that the glasses and ceramics are brittle materials; as a consequence of which, their handling and mechanical properties are not adequate for significant load-bearing applications. Therefore, a bioactive glass having the elastic modulus higher than that of bone is required for clinical applications. So, there was an increase in Young’s modulus from 76.363 to 78.894 GPa with increasing Sm2O3 content in the present base bioactive glass. Similar results were also obtained for shear and bulk moduli which showed improvement in the mechanical properties of bioactive glasses with varying ultrasonic velocities.

& & &

Funding information The authors gratefully acknowledge the support provided by the Central Instrument Facility Center (CIFC) of IIT (BHU), the Department of Ceramic Engineering, IIT (BHU), and the honorable Director of Indian Institute of Technology (BHU), Varanasi, 221005, India, for providing the necessary facilities for the present research work. The current work was supported by the Ministry of Human Resource Development, Government of India.

References 1.

Hench, L.L., Splinter, R.J., Allen, W.C., Greenlee, T.K.: Bonding mechanisms at the interface of ceramic prosthetic materials. J. Biomed. Mater. Res. Symp. 334, 117–141 (1971) 2. Hench, L.L., Splinter, R.J., Allen, W.C., Green-lee Jr, T.K.: Bonding mechanisms at the interface of ceramic prosthetic materials. J. Biomed. Mater. Res. 2(1), 117–141 (1971) 3. Beckham, C.A., Greenlee Jr., T.K., Crebo Jr., A.R.: Bone formation at a ceramic implant interface. Calcif. Tissue Res. 8(1), 165–171 (1971) 4. Greenlee Jr, T.K., Beckham, C.A., Crebo, A.R., Malmborg, J.C.: Glassceramic bone implants. J. Biomed. Mater. Res. 6(3), 235–244 (1972) 5. Srivastava, A.K., Pyare, R., Singh, S.P.: In vitro bioactivity and physical-mechanical properties of MnO2 substituted 45S5 bioactive glasses and glass-ceramics. J.f Biomat. Tissue Eng. 2(3), 249–258 (2012) 6. Bohner, M., Lemaitre, J.: Biomaterials. 30, 2175 (2009) 7. Hench, L.L., Wilson, J. (Eds.): An introduction to bioceramics (vol. 1). World Scientific (1993)

Author's personal copy 630

J Aust Ceram Soc (2018) 54:621–630

8.

system SiO2–Na2O–CaO–P2O5 and effect of gamma irradiation. Mater. Chem. Phys. 110(2–3), 352–362 (2008) Vyas, V.K., Kumar, S.A., Prasad, S., Singh, S.P., Pyare, R.: Bioactivity and mechanical behavior of cobalt oxide-doped bioactive glass. Bull. Mater. Sci. 38(4), 957–964 (2015) Vyas, V.K., Kumar, S.A., Prasad, S., Mohammad, E., Singh, S.P., Pyare, R.: Preparation and characterization of cobalt oxide doped 45S5 bioactive glass-ceramics. Innov. Corros. Mater. Sci. 5, 86–92 (2015) Vyas, V.K., Kumar, S.A., Singh, S.P., Pyare, R.: Effect of cobalt oxide substitution on mechanical behaviour and elastic properties of bioactive glass and glass-ceramics. Trans. Ind. Ceram. Soc. 75(1), 1–8 (2016) Vyas, V.K., Kumar, S.A., Singh, S.P., Pyare, R.: Effect of nickel oxide substitution on bioactivity and mechanical properties of bioactive glass. Bull. Mater. Sci. 39(5), 1355–1361 (2016) Vyas, V. K., Kumar, S. A., Singh, S. P., Pyare, R.: Destructive and non-destructive behavior of nickel oxide doped bioactive glass and glass-ceramic, J. Aust. Ceram. Soc. (2017) Kokubo, T., Takadama, H.: How useful is SBF in predicting in vivo bone bioactivity? Biomaterials. 27(15), 2907–2915 (2006) Arepalli, S.K., et al.: Influence of barium substitution on bioactivity, thermal and physic-mechanical properties of bioactive glass. Mater. Sci. Eng. C. 49, 549–559 (2015) Salman, S.M., Salama, S.N., Abo-Mosallam, H.A.: The role of strontium and potassium on crystallization and bioactivity of Na2O–CaO–P2O5–SiO2 glasses. Ceram. Int. 38(1), 55–63 (2012) Kansal, I., Goel, A., Tulyaganov, D.U., Santos, L.F., Ferreira, J.M.F.: Structure, surface reactivity and physicochemical degradation of fluoride-containing phospho-silicate glasses. J. Mater. Chem. 21(22), 8074–8084 (2011) Kim, C.Y., Clark, A.E., Hench, L.L.: Early stages of calciumphosphate layer formation in bioglasses. J. Non-Cryst. Solids. 113(2), 195–202 (1989) Tripathi, H., Kumar, A.S., Singh, S.P.: Preparation and characterization of Li2O–CaO–Al2O3–P2O5–SiO2 glasses as bioactive material. Bull. Mater. Sci. 39(2), 365–376 (2016) Hanan, H.B., Khaled, R.M., Gehan, T.E.-B.: Fabrication and characterization of bioactive glass (45S5)/titania biocomposites. Ceram. Int. 35, 1991–1997 (2009) Elfayoumi, M.A.K., Abdel-fattah, W.I., El-bassyouni, G.T.: Development of biomimetic coatings on Sm oxide doped ELB (Eu–Li–borate) glasses. Mater. Sci. Eng. C. 30(4), 509–517 (2010). https://doi.org/10.1016/j.msec.2010.01.012 Gandhi, Y., et al.: Influence of aluminum ions on fluorescent spectra and upconversion in codoped CaF2–Al2O3–P2O5–SiO2: Ho3+ and Er3+ glass system. J. Appl. Phys. 108, 023102 (2010)

Hench, L.L.: The story of Bioglass®. J. Mater. Sci. Mater. Med. 17(11), 967–978 (2006) 9. Hench, L.L., Thompson, I.: Twenty-first century challenges for biomaterials. J. R. Soc. Interface. 7(Suppl 4), S379–S391 (2010) 10. S. Jung: Borate based bioactive glass scaffolds for hard and soft tissue engineering Missouri University of Science and Technology (Ph.D. dissertation), (2010) 11. Jones, J.R., Ehrenfried, L.M., Saravanapavan, P., Hench, L.L.: Controlling ion release from bioactive glass foam scaffolds with antibacterial properties. J. Mater. Sci. Mater. Med. 17(11), 989– 996 (2006) 12. Diba, M., Tapia, F., Boccaccini, A.R.: Magnesium-containing bioactive glasses for biomedical applications. Int. J. Appl. Glas. Sci. 3(3), 221–253 (2012) 13. Deliormanlı, A.M.: Synthesis and characterization of cerium-and gallium-containing borate bioactive glass scaffolds for bone tissue engineering. J. Mater. Sci. Mater. Med. 26(2), 67 (2015) 14. Vyas, V.K., Kumar, A.S., Singh, S.P., Pyare, R., Ali, A., Prasad, S., Srivastava, E.P., Mallick, S.P.: Assessment of nickel oxide substituted bioactive glass-ceramic on in vitro bioactivity and mechanical properties. Boletín de la Sociedad Española de Cerámica y Vidrio. 69, 12 (2016) 15. Ershad, M., et al.: Synthesis and characterization of cerium- and lanthanum containing bioactive glass. Key Eng. Mater. 751, 617– 628 (2017) 16. Arunkumar, S., Marimuthu, K.: Concentration-effect of Sm3+ ions in B2O3–PbO–PbF2–Bi2O3–ZnO glasses–structural and luminescence investigations. J. Alloys Compd. 565, 104–114 (2013) 17. Selvi, S., Marimuthu, K., Muralidharan, G.: Effect of PbO on the B2O3–TeO2–P2O5–BaO–CdO–Sm2O3 glasses—structural and optical investigations. J. Non-Cryst. Solids. 461(2017), 35–46 (2017) 18. Yaru, N., Chunhua, L., et al.: Study on optical properties and structure of Sm203 doped boron-aluminosilicate glass. J Rare Earths Vol. 25, Suppl, p. 94 (2007) 19. Sayed, E., El-adawy, A., El-kheshkhany, N.: Effect of rare earth (Pr2O3, Nd2O3, Sm2O3, Eu2O3, Gd2O3, and Er2O3) on the acoustic properties of glass belonging to bismuth–borate system, 139, 108– 113 (2006) 20. Rygel, J.L., Pantano, C.G.: J. Non-Cryst. Solids. 355, 2622–2629 (2009) 21. Leonelli, C., Lusvardi, G., Malavasi, G., Menabue, L., Tonelli, M.: Synthesis and characterization of cerium-doped glasses and in vitro evaluation of bioactivity. J. Non-Cryst. Solids. 316, 198–216 (2003) 22. ElBatal, F.H., ElKheshen, A.: Preparation and characterization of some substituted bioglasses and their ceramic derivatives from the

23.

24.

25.

26.

27.

28. 29.

30.

31.

32.

33.

34.

35.

36.