You are on page 1of 6

Available online at www.sciencedirect.

com

CERAMICS
INTERNATIONAL
Ceramics International 41 (2015) S481–S486
www.elsevier.com/locate/ceramint

Dielectric, mechanical, and microstructural characterization


of HA–BST composites
S. Inthonga, T. Tunkasiria, G. Rujijanagula, K. Pengpata, C. Kruea-Inb, U. Intathac, S. Eitssayeama,n
a
Department of Physics and Materials science, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
b
Faculty of Science and Technology, Chiang Mai Rajabhat University, Chiang Mai 50300 Thailand
c
School of Science Mae Fah Luang University, Chiang Rai 57100, Thailand
Received 26 October 2014; accepted 14 March 2015
Available online 31 March 2015

Abstract

Hydroxyapatite (HA) is an excellent inorganic biomaterial and has various applications because its structure is similar to the inorganic matrix
of human bone and dental enamel. In this present study, the HA–(Ba0.7Sr0.3)TiO3; HA–BST composites were synthesized by a solid state reaction
technique. The phase, densification, microstructure, mechanical and dielectric properties of HA–BST composites sintered at 1350 1C were
investigated. The HA–BST composite showed an increase in densification, Vickers hardness, Young's modulus and fracture toughness with
maximum value as 3.42 g/cm3, 5.01 GPa, 161.70 GPa, and 1.48 MPa m1/2, respectively. The average grain size was decreased from 12.09 μm for
a reference HA sample to 1.14 μm for the 30 wt% BST in HA–BST composites. The dielectric constant as a function of temperature and
frequency tend to increase with increasing BST contents.
& 2015 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Keywords: B. Composites; C. Dielectric properties; C. Mechanical properties; D. Apatite

1. Introduction significantly accelerate or decelerate bone-like crystal growth on


the surface deposition in simulated body fluid (SBF), depending on
Many intensive studies have investigated the properties of electric polarization conditions, types and volume of surface charge
calcium phosphate with the objective of using in orthopedic and which are connected with dielectric properties. For improving
dental applications in the future. The investigation revealed several electrical and mechanical properties, the modified HA was focused
interesting properties such as good bioactivity, biocompatibility, to study by doping and synthesizing composites with HA such as
high wear resistance, osteoconductivity and the mineral component Fe-doping [12], Cd-doping [7] and HA–ZrO2 [13], HA–Sr [14],
similarity with nature mineral in bone and teeth [1–6]. However, HA–BaTiO3 [15], HA CaTiO3 [16]. Especially, HA–BaTiO3 and
calcium phosphate has many forms and phase such as hydro- HA–CaTiO3 have been presented as a good substrate for apatite
xypatite (HA), biphasic calcium phosphate (BCP), beta tricalcium growth with high osseointergration and osteoconction [15,17].
phosphate (β-TCP), alpha tricalcium phosphate (α-TCP), and Therefore, the properties of hydroxyapative–perovskite composites,
unsintered apatite (AP), which are dependent on fabrication process only non-toxic chemical compound, are highly potential for
conditions [7]. Recently, electrical properties such as dielectric, HA application due to their high polarizability and reasonable
ferroelectric and conductivity of HA have been of interest because biocompatibility.
these properties are linked to bioactivity property [3,8–11]. These BSTs (BaxSr1 xTiO3, 0oxo1) are ferroelectric materials
relations are interesting for many researches. Since Yamashita et al. that have high dielectric, low dielectric loss and spontaneous
[9] reported that the induced surface charge of polarized HA can polarization. The temperature of changing ferroelectric phase to
paraelectric phase was determined by the substitution of Ba2 þ by
n
Corresponding author. Sr2 þ in the A-site of perovskite structure (ABO3). Among BSTs,
E-mail address: sukum99@yahoo.com (S. Eitssayeam). Ba0.7Sr0.3TiO3 has a phase transition temperature of about 40 1C,

http://dx.doi.org/10.1016/j.ceramint.2015.03.205
0272-8842/& 2015 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
S482 S. Inthong et al. / Ceramics International 41 (2015) S481–S486

which is a high dielectric constant (εr  13,000) [18–20]. Th-


erefore, this material has high dielectric and still ferroelectric
phase at close to normal human body temperature.
Detailed literature survey reveals the new bone formation
has been found to be significantly improved on polarized
surface. For this significant context, many electroactive bio-
materials have been developed [15,17]. To our literature
reviews, a composite of HA–Ba0.7Sr0.3TiO3; HA–BST has
not yet been synthesized and their properties have not been
reported. Consequently, the effects of BST in HA–BST
composites were investigated and discussed compared with
other previous reports in this work.

2. Experimental

The nanocrystallinity of HA was performed by peroxide-


based route, with control pH at 9. Calcium nitrate solution
Fig. 1. XRD analysis of HA–BST composite as well as a reference HA.
(Ca(NO3)2  4H2O, Sigma-aldrich), ortho-phosphoric acid 85%
(Merck) and hydrogen peroxide solution 30% (Qrec) were
used as the stating reagent. During the chemical reaction, the
solution pH of the control was at a constant value of 9 by reference pattern spectra. The HA, α-TCP, and β-TCP phases
adding ammonium hydroxide solution 30% (J.T. Baker). The were detected in a reference HA and all HA–BST composi-
obtained powders of HA are in the nano-range ( 60 nm). The tes samples with reference numbers 09-0432, 09-0348, and 09-
BST powders were prepared by conventional solid state 0169 for HA, α-TCP, and β-TCP phases, respectively. It
processing. The reagent grades of BaCO3 (99.9%, Sigma- should be noted that the decomposition of HA was observed
aldrich), SrCO3 (99.9%, Sigma-aldrich), and TiO2 (99.9%, in reference HA and HA–BST composites. This result is in
Sigma-aldrich) were used as starting material. The mixed agreement with previous reports [21,22]. In general, HA phase
powers were calcined at 1150 1C for 4 h. The obtained powers transformation during thermal treatment consists of two
had perovskite structure and particle size about 0.9 μm. For the processes, which are dehydroxylation and decomposition.
preparation of the composite, the HA–xBST composites were Dehydroxylation to oxyhydroxyapatite proceeds at temperature
prepared by adding various amounts of BST contents with about 850–900 1C through the following equation [21]:
x=0, 5, 10, 15, 20, 25, and 30 wt% in HA. The mixed powers Ca10 ðPO4 Þ6 ðOHÞ2 -Ca10 ðPO4 Þ6 ðOHÞ2  2x Ox þ xH2 O ð1Þ
were ball milled in isopropanol for 24 h. This slurry was dried
and sieved to form homogeneous powders. These mixed For above 900 1C, the decomposition process of HA can be
powders were then granulated using polyvinyl alcohol (PVA) found according to the reaction in the following equation [22]:
3 wt% organic binder and formed into discs, typically 10 mm Ca10 ðPO4 Þ6 ðOHÞ2 -2Ca3 ðPO4 Þ2 þ Ca3 ðP2 O8 Þþ CaO þ H2 O
in diameter and 2 mm in thickness by cold uniaxial pressing.
ð2Þ
The resulting discs were sintered in air at a temperature of
1350 1C in a step of 100 1C/hour (with soaking time of 3 h). However, the perovskite structure of BST in HA–BST compo-
The polish of these ceramics was examined via an X-ray sites was shown at x4 5 wt%, according to JCPDS reference no.
diffractometer with Cu Kα radiation to check the phase 01-089-0274. The intensity of BST structure peaks is increased
formation. Density of samples was measured using the associated with BST contents in HA–BST composites. However,
Archimedes method. The surface morphology and elemental phases of BST with xr 5 wt% cannot be observed due to the
composition of the samples were then analyzed using a infinitesimal of BST contents were nearly a limit of this XRD
scanning electron microscope (SEM) equipped with an energy instrument. The morphology and composition of BST were
dispersive X-ray (EDX). Average grain size was determined analyzed by SEM and EDX as show in Fig. 3. The morphology
using the intercept method. Mechanical properties were studied form of SEM showed clearly two phases in the matrix of the HA–
on a polished surface using a micro-hardness tester. For BST surface, i.e., HA phase and BST phase. The phase of HA and
dielectric measurements, the ceramics were coated with silver BST was determined by EDX analysis. The spectra corresponding
paste electrodes on both sides of the pellets and the dielectric to HA showed a peak corresponding to that of phosphorus and
property was measured using an impedance analyzer. calcium as shown in x¼ 15 wt% BST. The EDX pattern of small
grain size in a matrix of HA–BST composites at x¼ 10 wt%
3. Results and discussion showed a peak corresponding to the major elements present in
composites (Ba, Sr, and Ti) as expected from BST grain.
Fig. 1 shows the XRD patterns of a reference HA, and In this work, reference HA and HA–BST composites
HA–BST composites ceramics. All studied samples were samples were sintered with different sintering temperatures
analyzed by comparison of peak location to JCPDS database for optimum sintering temperature condition. All samples were
S. Inthong et al. / Ceramics International 41 (2015) S481–S486 S483

maximum at x ¼ 25 wt% BST (161.70 GPa) and then


decreased. Similarly, the trend of fracture toughness is mostly
Young's modulus values. The highest value was recorded at
1.48 MPa m1/2 for x ¼ 25 wt% BST. It should be noted that
mechanical properties such as Vickers hardness, Young's
modulus and fracture toughness have greatly increased, which
is comparison a reference HA. The increasing of hardness and
fracture strength can be explained by grain size reduction and
matrix grain boundary reinforcement [27]. However, the
decrease of Young's modulus and fracture toughness at x ¼ 30
wt% BST may due to the increase of BST producing an
increasing inhomogeneous structure and pore in the HA matrix
Fig. 2. Bulk density and average grain size as a function of BST in the as shown in the SEM micrograph. However, the previous
HA–BST composite. reports [22,23] by other authors who studied a group of
calcium phosphate bioceramics demonstrated that the decom-
sintered at 1350 1C for 3 h, which produced the den- position of HA to TCP can lead to a significant decrease in
sest ceramic samples. The bulk density of the as-sintered mechanical strength. The grain sizes increased on increasing
HA–BST composites is shown in Fig. 2. It was observed that the temperature in the decomposition temperature range. For
the bulk density was increased from 2.98 g/cm3 to 3.42 g/cm3 our studies, the mechanical properties and grain sizes of the
for a reference HA to 30 wt% BST in HA–BST composites. HA–BST composites are mainly because of BST contents.
The increase of density in HA–BST may be because the These results are similarly found in other composites of HA as
theoretical density of the constituent compounds of BST mentioned in Sprio et al. and Silava et al. [24,25].
(5.69 g/cm3) [18] is higher than that of HA (3.16 g/cm3 by The variation in the dielectric constant of HA–BST compo-
assuming the absence of any reaction between the constituent sites under investigation was measured as a function of
phases), which is a normal phenomenon of density of temperature between 30 1C and 200 1C. Fig. 4(a) shows plots
composite ceramics. This result indicated that the BST of dielectric constant versus temperature for reference HA and
promoted the densification of HA–BST composites. HA–BST composites at 1 MHz. And the inset of Fig. 4(a)
The studied samples' surface morphology of HA–BST compo- shows temperature-dependent dielectric constant curve of BST
sites is shown in Fig. 3. The microstructure of all studied samples ceramics. It is revealed that dielectric constant increased on
showed dense ceramics without any cracks and fewer pores at the increasing the BST contents. Moreover, in the high frequency,
grain boundaries. SEM analysis revealed that BST additions the dielectric constant is weakly dependent on this temperature
produced a notable decrease in grain size of HA–BST composites. range. The characteristics of dielectric constant curve for the
An average value of grain size was determined by the intercept HA–BST composites agreed well with the work carried out by
method as shown in Fig. 2. The average grain sizes of the studied Dubey et al. [16,28]. This result may be due to the develop-
samples decreased from 12.09 μm for a reference HA sample to ment of interfacial polarization in the presence of HA and BST
1.14 μm for the x¼ 30 wt% BST in HA–BST composites. This phases in composites with different electrical conductivities.
result indicated that the addition of BST inhibited grain growth of Fig. 4(b) shows the dielectric constant as a function of
this composite. Moreover, both intergranular and intragranular frequency for reference HA and HA–BST composites. The
structures can be observed in HA–BST composites, which is dielectric constant exhibited frequency dispersion in which
predominated by the intergranular structure. This phenomenon is higher dielectric constants at lower frequencies are due to the
mostly found in the composites. During HA grain growth, residual presence of conductivity and different polarizations types in
pores were removed throughout the sinter of the HA matrix. this material. Fig. 4(c) shows a comparison of dielectric result
Simultaneously, BST particles were trapped in the grain boundaries with the theoretical estimations using parallel, series Wiener
(intergranular structure). This result is similar to the work reported bounds [28] and the logarithmic mixture rule [28], following
by Dubey et al. [16,17]. Therefore, the partial intergranular was the equation
increased on increasing the BST contents. It should be noted that
εc ¼ υHA εHA þ υBST εBST ð3Þ
the intergranular second phase small grain size acts to reduce the
grain size and pore and thus reinforces the grain boundaries. 1 υHA υBST
The mechanical properties of the studied samples were ¼ þ ð4Þ
εc εHA εBST
studied using Vickers hardness, Young's modulus and fracture
toughness. The mechanical properties of HA–BST compo-
log εc ¼ υHA log εHA þ υBST log εBST ð5Þ
sites are listed in Table 1. The results revealed that the
mechanical properties depended on the BST contents in the where εc is the dielectric constant of the HA–BST composites,
HA–BST composites. The Vickers hardness values increased υHA and υBST are the volume fractions of HA and BST, and εHA and
from 4.35 GPa for a reference HA sample to 5.01 GPa for εBST are the dielectric constants of HA and BST, respectively. The
x ¼ 30 wt% BST in HA–BST composites. Young's modulus of calculated dielectric constant was 12.58 for HA and 8102.47 for
samples increased with increasing BST content up to BST ceramics, respectively. The dielectric constant of the
S484 S. Inthong et al. / Ceramics International 41 (2015) S481–S486

Fig. 3. Selected SEM images and EDX analysis of the microstructures of the HA–BST composite.

composites follows closely the series Weiner bound, which is was studied by Dubey et al.[29]. They reported that the cell could
similar to the result of HA–CaTiO3 composite for CaTiO3 less than easily proliferate in the proximity of the composites of HA-40 wt%
60 wt% [16]. In addition, the relation of the biocompatibility BaTiO3 and HA-40 wt% CaTiO3 ceramics, for which the
(in vitro) and electrical properties of HA–BaTiO3 and HA–CaTiO3 room-temperature dielectric constants (1 kHz) were 18.7 and
S. Inthong et al. / Ceramics International 41 (2015) S481–S486 S485

22.7, respectively. It should be noted that the dielectric constants of were in the range of 2.98–3.42 g/cm3 and 12.09–1.14 μm,
HA–BST composites are nearly these dielectric values, which is respectively. The density of composites increased as a function
higher than the dielectric constant of dry human bone (9.2 at of BST contents while the average grain size decreased. In
1 kHz) [29]. addition, the mechanical and dielectric data indicated that BST
effectively improved both of these properties. The highest
4. Conclusions Vickers hardness and room temperature dielectric constants
(1 kHz) are 5.01 GPa and 32.87 for x ¼ 30 wt% BST, respec-
The composite of HA–BST ceramics was successfully tively. The highest values of Young's modulus  161.70 GPa
fabricated using conventional solid state processing with and fracture toughness  1.48 MPa m1/2, respectively, were
optimum sintering temperature at 1350 1C. This study inves- obtained for the sample with x¼ 25 wt% BST. Based on these
tigated the physical, mechanical and dielectric properties of results, this study suggests that BST had a significant effect on
these composites. All samples showed a decomposition of HA the properties of HA–BST composites.
in this sintering temperature. Density and grain size values
Table 1 Conflict of interest
The mechanical properties of HA–BST composites.
We declare that we do not have any commercial or associative
wt% BST Vickers hardness Young's modulus Fracture toughness interest that represents a conflict of interest in connection with the
(GPa) (GPa) (MPa m1/2)
work submitted.
0 4.35 30.73 0.59
5 4.56 59.10 0.66
10 4.62 80.80 0.91 Acknowledgment
15 4.69 78.67 1.04
20 4.79 110.50 1.28 The authors would like to thank Research and Researchers for
25 4.96 161.70 1.48 Industries (RRI) under the Thailand Research Fund (TRF), the
30 5.01 96.69 1.11
National Research University Project (NRU) under Thailand's
Human – 3–30 [25] 2–4 [26]
bone office of the Higher Education Commission (OHEC), Faculty of
Science, Chiang Mai University and Graduate School Chiang

Fig. 4. Temperature dependence of dielectric constant at 1 MHz (a) and frequency dependence of dielectric constant measured at 37.5 1C (b) as a function of BST.
A comparison of dielectric constant of HA–BST composites with theoretical estimation of Wiener bounds and logarithmic mixture rule at 30 1C and 1 MHz.
S486 S. Inthong et al. / Ceramics International 41 (2015) S481–S486

Mai University, Thailand, and National Research Council of [15] J.P. Gittings, C.R. Bowen, I.G. Turner, F. Baxter, J. Chaudhuri,
Thailand (NRCT) for financial support. Characterisation of ferroelectric-calcium phosphate composites and cera-
mics, J. Eur. Ceram. Soc. 27 (2007) 4187–4190.
[16] A.K. Dubey, P.K. Mallik, S. Kundu, B. Basu, Dielectric and electrical
References conductivity properties of multi-stage spark plasma sinter HA–CaTiO3
composites and comparison with conventionally sintered materials,
[1] M. Descamps, J.C. Hornez, A. Leriche, Manufacture of hydroxyapatite J. Eur. Ceram. Soc. 33 (2013) 3445–3453.
for medical applications, J. Eur. Ceram. Soc. 29 (2009) 369–375. [17] A.K. Dubey, G. Tripathi, B. Basu, Characterization of hydroxyapatite–
[2] J. Wang, L.L. Shaw, Nanocrystalline hydroxyapatite with simultaneous perovskite (CaTiO3) composites: phase evaluation and cellular response,
enhancements in hardness and toughness, Biomaterials 30 (2009) J. Biomed. Mater. Res. Part B: Appl. Biomater. 95 (2010) 320–329.
6565–6572. [18] Y.C. Liou, C.T. Wu, Synthesis and diffused phase transition of
[3] J.P. Gittings, C.R. Bowen, A.C.E. Dent, I.G. Turner, F.R. Baxter, Ba0.7Sr0.3TiO3 ceramics by a reaction-sintering process, Ceram. Int. 34
J.B. Chaudhuri, Electrical characterization of hydroxyapatite-based bio- (2008) 517–522.
ceramics, Acta Biomater. 5 (2009) 743–754. [19] S.W. Kim, H.I. Choi, M.H. Lee, J.S. Park, D.J. Kim, D. Do, M.H. Kim,
[4] B. He, S. Huang, J. Jing, Y. Hao, Measurement of hydroxyapatite density T.K. Song, W.J. Kim, Electrical properties and phase of BaTiO3–SrTiO3
and Knoop hardness in sound human enamel and a correlational analysis solid solution, Ceram. Int. 39 (2013) S487–S490.
between them, Arch. Oral Biol. 55 (2010) 134–141. [20] X.Q. Liu, T.T. Chen, M.S. Fu, Y.J. Wu, X.M. Chen, Electrocaloric
[5] M. Qidwai, M.A. Sheraz, S. Ahmed, A.A. Alkhuraif, I.U. Rehman, effects in spark plasma sintered Ba0.7Sr0.3TiO3-based ceramics: effects of
Preparation and characterization of bioactive composites and fibers for domain sizes and phase constitution, Ceram. Int. 40 (2014) 11269–11276.
dental application, Dent. Mater. 30 (2014) e253–e263. [21] S. Ramesh, C.Y. Tan, W.H. Yeo, R. Tolouei, M. Amiriyan, I. Sopyan,
[6] V.V.A. Thampi, M. Prabhu, K. Kavitha, P. Manivasakan, P. Prabu, W.D. Teng, Effects of bismuth oxide on the sinterability of hydroxya-
V. Rajendran, S. Shankar, P. Kulandaivelu, Hydroxyapatite, alumina/ patite, Ceram. Int. 37 (2011) 599–606.
zirconia, and nanobioactive glass cement for tooth-restoring application, [22] K.T. Chu, S.F. Ou, S.Y. Chen, S.Y. Chiou, H.H. Chou, K.L. Ou,
Ceram. Int. 40 (2014) 14355–14365. Research of phase transformation induced biodegradable properties on
[7] K.K. Bamzai, S. Suri, V. Singh, Synthesis, characterization, thermal and hydroxyapatite and tricalcium phosphate based bioceramic, Ceram. Int.
dielectric properties of pure and cadmium doped calcium hydrogen 39 (2013) 1455–1462.
phosphate, Mater. Chem. Phys. 135 (2012) 158–167. [23] K.C.B. Yeong, J. Wang, S.C. Ng, Mechanochemical synthesis of
[8] T.P. Hoepfner, E.D. Case, The porosity dependence of the dielectric nanocrystalline hydroxyapatite, Biomaterials 22 (2001) 2705–2712.
constant for sintered hydroxyapatite, J. Biomed. Mater. Res. 60 (2002) [24] S. Sprio, S. Guicciardi, M. Dapporto, C. Melandri, A. Tampieri,
643–650. Synthesis and mechanical behavior of β-tricalcium phosphate/titania
[9] K. Yamashita, N. Oikawa, T. Umeqaki, Acceleration and deceleration of composites addressed to regeneration of long bone segment, J. Mech.
bone-like crysal growth on ceramic hydroxyapatite by electric poling, Behav. Biomed. Mater. 17 (2013) 1–10.
Chem. Mater. 8 (1996) 2697–2700. [25] V.V. Silva, F.S. Lameiras, R.Z. Domingues, Microstructural and mechan-
[10] N.C. Teng, S. Nakamura, Y. Takagi, Y. Yamashita, M. Ohgaki, ical study of zirconia–hydroxyapatite (ZH) composite ceramics for
K. Yamashita, A new approach to enhancement of bone formation by biomedical applications, Compos, Sci. Technol. 61 (2001) 301–310.
electrically polarized hydroxyapatite, J. Dent. Res. 80 (2001) 1925–1929. [26] T.L. Norman, D. Vashishth, D.B. Burr, Fracture toughness of human
[11] S. Agrawal, R. Guo, D.K. Agrawal, A.S. Bhalla, R.R. Neurgaonkar, bone under tension, J. Biomech. 28 (3) (1995) 309–320.
C.B. Murray, Dielectric tunability of BST:MgO composites prepared by [27] R.W. Rice, Correlation of machining-grain size effects on tensile strength
using nano-particles, Ferroelectr. Lett. 31 (2004) 149–156. with tensile strength-grain size behavior, J. Am. Ceram. Soc. 76 (1993)
[12] O. Kaygili, S.V. Dorozhkin, T. Ates, A.A. Al-Ghamdi, F. Yakuphanoglu, 1068–1070.
Dielectric properties of Fe doped hydroxyapatite prepared by sol–gel [28] A.K. Dubey, B. Basu, K. Balani, R. Guo, A.S. Bhalla, Dielectric and
method, Ceram. Int. 40 (2014) 9395–9402. pyroelectric properties of HAp–BaTiO3 Composites, Ferroelectrics 423
[13] M. Sh Kalil, H.H. Beheri, W.I.A. Fattah, Structural and electrical (2011) 63–76.
properties of zirconia/hydroxyapatite porous composites, Ceram. Int. 28 [29] A.K. Dubey, B. Basu, K. Balani, R. Guo, A.S. Bhalla, Multifunctionality
(2002) 451–458. of perovskites BaTiO3 and CaTiO3 in a composite with hydroxyapatite as
[14] J. Abert, C. Bergmann, H. Fischer, Wet chemical synthesis of strontium- orthopedic implant materials, Integr. Ferroelectr. 131 (2011) 119–126.
substituted hydroxyapatite and its influence on the mechanical and
biological properties, Ceram. Int. 40 (2014) 9195–9203.

You might also like