DOI:
10.1039/C4RA02441A
(Paper)
RSC Adv., 2014,
4, 29946-29956
Quantitative analysis of the structural stability and degradation ability of hydroxyapatite and zirconia composites synthesized in situ
Received
20th March 2014
, Accepted 29th May 2014
First published on 6th June 2014
Abstract
In situ synthesis has been attempted in the present study to form hydroxyapatite [HAP, Ca10(PO4)6(OH)2] and zirconia [ZrO2] composite powders. The synthesis has been attempted using an aqueous precipitation technique in which alterations have been made to the starting Ca/P molar ratio of calcium phosphate suspensions together with five different concentrations of ZrO2 precursors. An additive in the form of yttria to stabilize tetragonal ZrO2 (t-ZrO2) has not been used in the present investigation. The precipitated powders resulting were heat treated at different temperatures before determination of the phase stability and degradation ability of the HAP–ZrO2 composites. The powders were characterized by X-ray diffraction, Raman spectroscopy and quantitative analysis by Rietveld refinement. The results from the investigation confirmed the formation of HAP–ZrO2 composites at 900 °C. The Ca/P precursors with molar ratios of 1.67 and 1.69 have yielded the formation of β-tricalcium phosphate [β-TCP, β-Ca3(PO4)2] with a content of less than 5 wt% together with HAP and t-ZrO2 at 900 °C. The Ca/P precursors with a molar ratio of 1.68 have yielded only HAP and t-ZrO2 composite powders at 900 °C. The increased addition of ZrO2 precursors had resulted in the enhanced formation of t-ZrO2 with a simultaneous reduction in the amount of the HAP phase formed. However, heat treatment at 1000 °C and 1100 °C has initiated the degradation of HAP and t-ZrO2 composite powders resulting in the formation of monoclinic ZrO2 (m-ZrO2). It is surmised from the present results that there is active involvement of the ZrO2 content in the degradation behaviour of the HAP–ZrO2 composite powders.
1. Introduction
The stoichiometric hydroxyapatite [HAP, Ca10(PO4)6(OH)2] and several ionic substitutions in HAP has found many applications in the treatment of bone- and tooth-related defects for the past three decades because of their excellent biocompatibility and chemical similarities with natural bone and the minerals in teeth.1–6 The inferior mechanical features of synthetic HAP in comparison with the biological bone have restricted its applications to the form of particulates or granules to fill either bone- or tooth-related defects. Several attempts have been made to recognize synthetic HAP as a suitable substitute in the replacement of hard tissues where the demands for high mechanical properties are necessary. Therefore, HAP has been used in the form of a reinforcement either with metals such as titanium alloy and stainless steels or with bioinert ceramics such as alumina (Al2O3), zirconia (ZrO2) and titania (TiO2). The HAP coatings on metallic implants have serious limitations such as metallic corrosion, and weak interfacial bonding between the metal and the ceramic, which is expected to result in the dissolution of the ceramic.7–10 The plasma spray technique used to develop a HAP coating on metallic implants resulted in strong adhesion between the ceramic and the metallic alloy; however, such coatings had resulted in the phase decomposition of stoichiometric HAP to yield secondary phases. Al2O3 has been known for its features such as high mechanical stability, corrosion resistance and high wear resistance.11 Generally the phase stabilization of Al2O3 occurs at a higher temperature whereas the phase decomposition of HAP starts to occur beyond 1200 °C and this leads to the dehydroxylation and conversion of HAP to tetracalcium phosphate [Ca4(PO4)2O] and α-tricalcium phosphate [α-TCP, α-Ca3(PO4)2]. This mismatch in the thermal stability of HAP and Al2O3 composites has led to difficulties in the preservation of their respective phases in the HAP/Al2O3 composite and some researchers have also reported on the chemical interaction between HAP and Al2O3 to form a CaAl2O4 phase.12,13 The investigation of the reinforcement of TiO2 with HAP has been mainly based on the fabrication of HAP/TiO2 composite coatings on titanium metal. The reason for the development of such composite coatings on titanium metal is mainly based on the fact that the reinforcement of bioinert TiO2 in the HAP matrix would substantially improve the mechanical reliability of HAP, thereby leading to the increased stability of the coatings.14,15 The thermal decomposition of HAP/TiO2 composites to form calcium oxide (CaO), calcium titanate and the combination of rutile TiO2 and anatase TiO2 have also been reported.16 ZrO2 has been known to possess better mechanical features when compared to Al2O3
17 and their comparative investigations have concluded that ZrO2 has shown better resistance to fracture compared to Al2O3.18
It is well known that ZrO2 exists in three polymorphs: monoclinic (stable from room temperature to 1170 °C with P21/c symmetry), tetragonal (stable between 1170 °C to 2370 °C with P42/nmc symmetry) and cubic (stable above 2370 °C with Fm3m symmetry).19 Among these polymorphs, tetragonal ZrO2 (t-ZrO2) has been found to be suitable for hard tissue replacements because it has favourable features such as wear resistance, chemical resistance, high fracture toughness and hardness. Many attempts have been made to combine the advantages of biocompatible HAP and mechanically stable ZrO2 to form a composite which will be suitable for hard tissue replacements. The reported methods for the fabrication of HAP–ZrO2 composites include the mechanical mixing of either a commercially procured or a synthesized individual HAP and ZrO2 particles to form a composite.20–25 The conventional heat treatment of the mechanically mixed individual powder particles of HAP and ZrO2 give the formation secondary phases such as CaZrO3, α-Ca3(PO4)2, Ca4(PO4)2O, and CaO, in addition to the presence of HAP and ZrO2 phases. Thus, the presence of additional phases has led to difficulties in the precise determination of the quantitative content of HAP and ZrO2 in the composites. Some of the studies have also reported on the application of a spark plasma sintering process for mechanically mixing the individual HAP and yttria (Y2O3)-stabilized ZrO2 (y-ZrO2) particles to yield a composite consisting only of the phases of HAP and ZrO2 particles.26–28 Such HAP–ZrO2 composites have resulted in a variable mechanical strength that was solely dependent on the amount y-ZrO2 added to the composites. Thus, much of the reported research on the development and fabrication of HAP–ZrO2 composites has only dealt with the mechanical mixing of the separately synthesized HAP and ZrO2 particles to form composites. The drawbacks in such studies are the difference in the grain growth of HAP and ZrO2 particles during sintering, non-uniformity in the mixing level of the individual HAP and ZrO2 particles throughout the composite matrix and lack of precise determination of the quantitative content of individual particles in the matrix. Thus, the present study has been focussed on the in situ synthesis of five different HAP–ZrO2 composites starting from the solution precursors and followed by the application of the Rietveld refinement technique for the quantitative determination of the resulting individual HAP and ZrO2 particles in the composites.
2. Materials and methods
2.1 Powder synthesis
An aqueous co-precipitation technique was used in the present investigation to synthesise different concentrations of HAP–ZrO2 ceramic composite powders. Calcium nitrate tetrahydrate [Ca(NO3)2·4H2O, HiMedia, India], ammonium di-hydrogen orthophosphate [NH4H2PO4, HiMedia, India] and zirconium oxychloride octahydrate [ZrOCl2·8H2O, HiMedia, India] were used as precursors for Ca2+, PO43− and ZrO2. For the synthesis of the different HAP–ZrO2 composite powders, stock solutions of appropriate molar concentrations of Ca(NO3)2 and NH4H2PO4 were prepared separately in Millipore water. The molar concentrations of Ca(NO3)2 and NH4H2PO4 precursor solutions were maintained for all the five different concentrations in order to ensure the stoichiometric Ca/P molar ratio of 1.67 for all the five different compositions. Five different concentrations of a ZrOCl2·8H2O stock solution ranging from 0.1 M to 0.5 M were prepared separately in Millipore water. In addition to the Ca/P molar ratio of 1.67, two more Ca/P ratios of 1.68 and 1.69 were also synthesized in situ with the ZrO2 precursors to form a wide range of composite powders. A total of 15 different concentrations were prepared and their respective sample codes for all the 15 different compositions are indicated in Table 1.
Table 1 Molar concentrations of the precursors used in the synthesis of HAP–ZrO2 composite powders
No. |
Sample code |
Molar concentrations of precursors |
Ca/P molar ratio |
Ca/Zr molar ratio |
Ca(NO3)2·4H2O |
NH4H2PO4 |
ZrOCl2·8H2O |
1. |
1.67CaP_1xZrO2 |
0.5 M |
0.2994 M |
0.1 |
1.67 |
5.00 |
2. |
1.67CaP_2xZrO2 |
0.5 M |
0.2994 M |
0.2 |
1.67 |
2.50 |
3. |
1.67CaP_3xZrO2 |
0.5 M |
0.2994 M |
0.3 |
1.67 |
1.67 |
4. |
1.67CaP_4xZrO2 |
0.5 M |
0.2994 M |
0.4 |
1.67 |
1.25 |
5. |
1.67CaP_5xZrO2 |
0.5 M |
0.2994 M |
0.5 |
1.67 |
1.00 |
6. |
1.68CaP_1xZrO2 |
0.5 M |
0.2976 M |
0.1 |
1.68 |
5.00 |
7. |
1.68CaP_2xZrO2 |
0.5 M |
0.2976 M |
0.2 |
1.68 |
2.50 |
8. |
1.68CaP_3xZrO2 |
0.5 M |
0.2976 M |
0.3 |
1.68 |
1.67 |
9. |
1.68CaP_4xZrO2 |
0.5 M |
0.2976 M |
0.4 |
1.68 |
1.25 |
10. |
1.68CaP_5xZrO2 |
0.5 M |
0.2976 M |
0.5 |
1.68 |
1.00 |
11. |
1.69CaP_1xZrO2 |
0.5 M |
0.2959 M |
0.1 |
1.69 |
5.00 |
12. |
1.69CaP_2xZrO2 |
0.5 M |
0.2959 M |
0.2 |
1.69 |
2.50 |
13. |
1.69CaP_3xZrO2 |
0.5 M |
0.2959 M |
0.3 |
1.69 |
1.67 |
14. |
1.69CaP_4xZrO2 |
0.5 M |
0.2959 M |
0.4 |
1.69 |
1.25 |
15. |
1.69CaP_5xZrO2 |
0.5 M |
0.2959 M |
0.5 |
1.69 |
1.00 |
As a preliminary step in the synthesis, the stock solution of NH4H2PO4 was added dropwise to the continuously stirred Ca(NO3)2 solution. The temperature of the stirred suspension was maintained at 80 °C throughout the reaction. After the completion of the NH4H2PO4 addition, the pH of the suspension was increased from the initial value of ∼4.00 to 10.00 by the addition of concentrated NH4OH solution. The resulting mixture was allowed to stir for another 30 minutes to ensure the formation of a calcium phosphate precursor. Then, the separately prepared stock solution of ZrOCl2 was added dropwise to the continuously stirred calcium phosphate mixture and the resulting mixtures containing the precursors of Ca2+, PO43− and ZrO2 were continuously stirred for two hours. The resulting mixture was then transferred to a hot air oven and dried at 120 °C for 48 hours. The dried precipitates obtained were ground into fine powder with a pestle and mortar and this powder was considered to be the as-prepared composite powder. Pure HAP and pure ZrO2 have also been prepared simultaneously by adapting a similar aqueous precipitation technique for comparative purposes.
2.2 Characterization of powders
All the finely ground as prepared powder samples were heat treated at various temperatures to investigate the ability of the formation of a varied range of HAP–ZrO2 composites. The powder samples were subjected to heat treatment at a rate of 5 °C per minute over various temperature ranges with a soaking time of two hours in a muffle furnace (MATRI-MC 2265 A, India) and subjected to X-ray diffraction (XRD) to observe any change in the phase behaviour and crystallinity. The phase purity and composition of all the composite powders were determined using a powder X-ray diffractometer (RIGAKU, Ultima IV, Japan). XRD studies for all the powders were carried out with Cu Kα radiation (λ = 1.5406 nm) produced at 40 kV and 30 mA to scan the diffraction angles (2θ) between 10° and 70° with a step size of 0.02° 2θ per second. Phase determinations were made using standard International Centre for Diffraction Data (ICDD) card no. 00-009-0432 for HAP, 01-079-1765 for t-ZrO2, 01-071-4810 for cubic ZrO2 (c-ZrO2), 01-083-0944 for monoclinic ZrO2 (m-ZrO2) and 00-009-0169 for β-Ca3(PO4)2.
The vibrational modes for HAP–ZrO2 composite powders were determined by using Raman spectroscopy. All the vibrational modes of HAP–ZrO2 composite powders after heat treatment at 900 °C, 1100 °C and 1100 °C were determined by using the back scattering geometry of a confocal Raman microscope (Renishaw, UK). All the powder samples were excited at a wavelength of 785 nm by a semiconductor diode laser (0.5% of power) with a data acquisition time of 30 seconds. The powder samples were analysed in the spectral range of 100 to 1200 cm−1.
2.3 Quantitative analysis by Rietveld refinement
Rietveld analysis was used to provide quantitative information about the phases present in the sample. In the present study, quantitative phase analysis was performed for stoichiometric HAP, ZrO2 and five different HAP–ZrO2 composites samples by using the Rietveld method using a GSAS-EXPGUI software package.29,30 For the analysis by the Rietveld refinement, an average of three scans were recorded for each powder samples with Cu Kα radiation (λ = 1.5406 nm) produced at 40 kV and 30 mA to scan the diffraction angles (2θ) between 10° and 70° with a step size of 0.02° 2θ per second using an X-ray powder diffractometer (RIGAKU, Ultima IV, Japan). Quantitative refinement was carried out for all the five different HAP–ZrO2 composites, and the stoichiometric HAP and ZrO2 were determined for powder samples heat treated at 900 °C, 1000 °C and 1100 °C. All the standard crystallographic information was obtained from the American mineralogist crystal structure database. The standard crystallographic data for the refinement of HAP, t-ZrO2, β-Ca3(PO4)2 and m-ZrO2 were obtained from Wilson et al.,31 Howard et al.,32 Yashima et al.33 and Smith and Newkirk,34 respectively. In structural refinement, numerous cycles were run to achieve the quantitative analysis, weight fraction and bond length of the composite powders. In the first step of the refinement, all the structural parameters were fixed to values from the literature. Then, during the successive refinement cycles, numerous parameters were allowed to vary according to the relative weight of the observed phases. The following refinement sequence has been used as a standard for all the structures: scale factor, zero shift, background as Chebyshev polynomial of the fifth grade, peak profile, and lattice parameters. Fittings were performed using pseudo-Voigt peak profile functions and a preferred orientation along [001] was accounted for with the Marsh model. The fractional coordinates, isotropic temperature and atomic parameters were employed during refinement.
3. Results and discussion
3.1 Formation of HAP–ZrO2 composite powders
The as-prepared powders were subjected to heat treatment at various temperatures in order to analyse the phase transformation behaviour. The XRD patterns of the as prepared powders (Fig. 1 and 2) were characterized by broad diffraction bands and no significant information could be obtained because of the excessive amorphous content present in the samples which had mainly arisen from the precursor impurities (NH4+, Cl−, NO32− and adsorbed H2O) of the aqueous precipitation technique. Fig. 1 and 2 present the corresponding XRD patterns of the 1.68CaP_1xZrO2 and 1.68CaP_3xZrO2 samples after heat treatment at different temperatures. The heat treatment at 500 °C had resulted in the emergence of diffraction patterns characteristic of the HAP phase and no peaks were present for the ZrO2 phase. The heat treatment at 700 °C had resulted in the formation of well-defined diffraction patterns that corresponded to the HAP phase and the formation of the ZrO2 phase could be identified by the appearance of a small peak at the 2θ angle of 30°. Successive heat treatments at higher temperatures ranging from 800 °C to 1100 °C had resulted in the enhancement of diffraction patterns that corresponded to the presence of both HAP and ZrO2 phases. However, the heat treatment at 1000 °C and 1100 °C had resulted in the formation of very small intensity diffraction patterns, which were characteristic of the β-Ca3(PO4)2 phase, which has not been identified in the sample that was heat treated at 900 °C. The characteristic peaks for the presence of m-ZrO2 phase were detected at 1100 °C for all the samples, despite the varied ZrO2 precursor concentration added during the synthesis and also the varied Ca/P molar concentration maintained during the synthesis. It was observed from a previous investigation by the current authors35 that pure ZrO2 without the addition of any precursors had resulted in the formation of mixtures comprising c-ZrO2 and m-ZrO2 phase at 500 °C and heat treatment at elevated temperatures (700 °C, 900 °C and 1100 °C) had resulted in the enhanced presence of m-ZrO2 with the corresponding presence of a c-ZrO2 phase at the reduced level. It should be noted that the presence of the t-ZrO2 phase was not detected at any of the heat treatment temperatures used in the investigation of the pure ZrO2 system. The results from the present investigation have shown that the presence of the HAP phase had restricted the formation of the c-ZrO2 and m-ZrO2 phases at the temperatures studied until 1000 °C was reached and simultaneously it had stabilized the t-ZrO2 phase.
 |
| Fig. 1 XRD patterns of the 1.68CaP_1xZrO2 sample after heat treatment at different temperatures. | |
 |
| Fig. 2 XRD patterns of the 1.68CaP_3xZrO2 sample after heat treatment at different temperatures. | |
The diffraction patterns presented in Fig. 1 and 2 show only the presence of pure HAP and t-ZrO2 phases at 900 °C, which resulted in the formation of HAP–ZrO2 composite powders when the Ca/P molar ratio was maintained at 1.68 during the synthesis. The diffraction patterns of the composite powders had initiated the formation of β-Ca3(PO4)2and m-ZrO2 after heat treatment at 1000 °C and 1100 °C. Both the HAP phase and the t-ZrO2 had shown excellent matches with the standard ICDD diffraction patterns of card nos 00-009-0432 for HAP and 01-079-1765 for t-ZrO2. The increased level of ZrO2 precursor added during the synthesis had resulted in a corresponding enhancement in the intensity of the diffraction patterns observed for t-ZrO2 phase with the simultaneous reduction in the intensity of diffraction patterns observed for the HAP phase. A similar trend could be noticed in the XRD patterns presented in Fig. 1 and 2, with, however, the formation of an additional phase in the form of β-Ca3(PO4)2 and m-ZrO2 phases, which had confirmed the decomposition of the HAP phase at 1000 °C. A previous study had illustrated the comparative decomposition mechanism of mechanically mixed HAP–ZrO2 composite after heat treatment in air and hot isostatic pressing. According to work reported by Adolfsson et al.,26 the heat treatment in air at less than 950 °C had resulted in the decomposition of the HAP–ZrO2 composite by the reactions given next:
|
Ca10(PO4)6(OH)2 → Ca10(PO4)6(OH)2−2xOx + xH2OCa10(PO4)6(OH)2−2xOx + yZrO2(tetragonal) → Ca3(PO4)2 + CaO(ZrO2(cubic))y + (1 − x)H2O
| (1) |
However, the pressure-assisted sintering, hot isostatic pressing or spark plasma sintering of the HAP–ZrO2 composite samples have prevented the decomposition of HAP–ZrO2 composites until 1200 °C is reached.23,36,37 It should be noted that in the research reported above, the yttria precursor has been used as a stabilizing agent for the preservation of the t-ZrO2 phase. In the present investigation, without the usage of a yttria precursor, the heat treatment at 900 °C did not cause the decomposition of the HAP–ZrO2 composite with the maintained precursor molar ratio of Ca/P being 1.68, whereas at 1000 °C and 1100 °C, the formation of the m-ZrO2 and the β-Ca3(PO4)2 phase has been detected in all the composite powders. The decomposition of HAP in the HAP–ZrO2 composite system during heat treatment in air has been attributed to vacancy formation during the release of structural water.26 In the present investigation, HAP–ZrO2 composite powders have been formed by an in situ precipitation technique and during the heat treatment process, the gradual elimination of volatile impurities (NH4+, Cl−, NO32− and adsorbed H2O) that generally arise during the aqueous precipitation technique has to occur, followed by the release of structural water. This elimination of volatile impurities has been evident from the XRD patterns (Fig. 1 and 2) through the conversion of observed broad diffraction patterns at lower temperatures (500 °C) to the sharp diffraction patterns at higher temperatures (900 °C). The elimination of structural water is initiated only after the removal of volatile impurities has been completed. According to the present results, the decomposition of HAP has started to occur by the formation of m-ZrO2 and β-Ca3(PO4)2 at both 1000 °C and 1100 °C. The XRD patterns for all the powder samples after heat treatment at 1200 °C had resulted in the formation of α-TCP thus signifying the further degradation of the composites. However, the XRD results have not given the exact quantitative information for the yield of HAP, ZrO2 and β-Ca3(PO4)2 phases in all the compositions investigated.
3.2 Structural analysis of the HAP–ZrO2 composite powders
The results from the XRD analysis have confirmed the formation of HAP–ZrO2 composites at 900 °C and heat treatment at 1000 °C has resulted in the formation of m-ZrO2 and β-Ca3(PO4)2 phases. However, the structural information of the phases formed could not be specified from the XRD analysis and thus, a quantitative analysis in the form of a Rietveld refinement has been employed for a precise analysis of the structural and compositional behaviour of the HAP–ZrO2 composite powders. It has been already reported by Vasanthavel et al.35 and Nandha Kumar et al.38 that the Rietveld refinement technique can be employed for the quantitative analysis of the bioceramics. Fig. 3 presents the Rietveld analysis pattern of the powder diffraction data of the HAP and HAP–ZrO2 composite powders after heat treatment. Table 2 presents the data of Rwp, Rp, χ2, and RBragg obtained from the refinement of different HAP–ZrO2 composites after heat treatment at three different temperatures (900 °C, 1000 °C and 1100 °C) and these values showcase the refinement standards executed in the present investigation. The powder samples heat treated at 1200 °C have not been subjected to a refinement study because of the allotropic phase conversion behaviour of β-TCP to α-TCP that generally occurs at 1180 °C which is followed by the reconversion of α-TCP to β-TCP during cooling to the room temperature. As can be seen from the Table 2, the goodness of fit (GoF) values were relatively low, i.e., all were approximately less than 2%, which is considered as acceptable according to basic principle of a GoF less than 4% and a Rwp of less than 20%.39 The refined lattice parameters and unit cell volume of the stoichiometric HAP and the HAP phase present in all the HAP–ZrO2 composite powders confirms the hexagonal setting that crystallizes in P63/m (176) space group at all the heat treatment temperatures of 900 °C, 1000 °C and 1100 °C (Table 3). However, the heat treatment at 1000 °C had resulted in the additional formation of a β-Ca3(PO4)2 phase, which possesses a rhombohedral structure that tends to crystallize in the R3c (167) space group. In a similar manner, the refined lattice data of the ZrO2 phase in all the HAP–ZrO2 composite powders confirms the tetragonal setting that crystallizes in the P42/nmc (137) space group at the investigated temperatures of 900 °C, 1000 °C and 1100 °C. The formation of an additional ZrO2 phase at 1000 °C has confirmed its monoclinic setting in the P21/c14 space group. Despite the different amounts of ZrOCl2 precursor that were added during the synthesis, the a-axis and c-axis lattice parameter values of the HAP phase have shown a uniform trend in all the HAP–ZrO2 composite powders at 900 °C, 1000 °C and 1100 °C, whereas the lattice parameter values of t-ZrO2 have also indicated a uniform trend in all the HAP–ZrO2 composite powders at 900 °C, 1000 °C and 1100 °C. The lattice data obtained for the HAP phase and the t-ZrO2 phase over various temperatures indicates the fact that the individual phases are unique in the composite powders.
 |
| Fig. 3 Rietveld analysis pattern of powder diffraction data of the HAP and HAP–ZrO2 composite powders after heat treatment.(a) Stoichiometric HAP at 900 °C, (b) 1.68CaP_5xZrO2 at 1000 °C, (c) 1.68CaP_5xZrO2 at 1100 °C *The solid lines are calculated intensities and the marked ones are the intensities observed. The difference between the observed and calculated intensities is plotted below the profile. | |
Table 2 Rietveld agreement factors obtained for the different HAP–ZrO2 composite powders after heat treatment at different temperatures
Sample code |
900 °C |
1000 °C |
1100 °C |
cRpa |
cRwpa |
χ2b |
RBragg |
cRpa |
cRwpa |
χ2b |
RBragg |
cRpa |
cRwpa |
χ2b |
RBragg |
cRp and cRwp represent the Rietveld agreement factors. χ2 is a goodness of fit (GoF) indicator defined by (Rwp/Rexp)2 in which Rexp is the expected weighted profile factor. RBragg demonstrates the agreement between the observed and computed reflection during refinement. |
1.67CaP_1xZrO2 |
09.80 |
7.42 |
1.542 |
05.45 |
09.57 |
07.30 |
1.497 |
05.82 |
09.63 |
7.42 |
1.616 |
09.98 |
1.67CaP_2xZrO2 |
09.30 |
6.76 |
1.576 |
09.10 |
08.38 |
06.45 |
1.319 |
05.80 |
14.03 |
10.5 |
3.809 |
11.99 |
1.67CaP_3xZrO2 |
09.21 |
6.81 |
1.681 |
11.73 |
09.56 |
07.27 |
1.836 |
08.19 |
07.25 |
5.65 |
1.063 |
03.82 |
1.67CaP_4xZrO2 |
09.44 |
6.94 |
1.843 |
10.53 |
07.99 |
06.07 |
1.349 |
07.53 |
07.92 |
6.13 |
1.325 |
04.94 |
1.67CaP_5xZrO2 |
08.80 |
6.55 |
1.683 |
08.15 |
08.08 |
06.03 |
1.442 |
04.89 |
07.56 |
5.83 |
1.284 |
04.96 |
1.68CaP_1xZrO2 |
11.03 |
8.22 |
1.970 |
05.75 |
09.71 |
07.43 |
1.475 |
05.84 |
11.64 |
9.03 |
2.166 |
09.52 |
1.68CaP_2xZrO2 |
10.92 |
8.22 |
2.145 |
08.95 |
09.92 |
07.84 |
1.828 |
06.74 |
10.85 |
8.38 |
2.258 |
10.15 |
1.68CaP_3xZrO2 |
11.13 |
8.23 |
2.510 |
06.95 |
09.95 |
07.79 |
1.893 |
08.48 |
09.27 |
7.20 |
1.710 |
07.41 |
1.68CaP_4xZrO2 |
10.53 |
7.80 |
2.345 |
06.13 |
08.62 |
06.79 |
1.509 |
07.09 |
08.96 |
6.94 |
1.703 |
08.57 |
1.68CaP_5xZrO2 |
11.05 |
7.97 |
2.717 |
08.76 |
06.86 |
05.19 |
1.025 |
04.22 |
07.49 |
5.70 |
1.264 |
03.62 |
1.69CaP_1xZrO2 |
10.36 |
7.97 |
1.753 |
06.90 |
11.20 |
08.57 |
1.945 |
07.74 |
10.71 |
8.28 |
1.863 |
08.36 |
1.69CaP_2xZrO2 |
09.43 |
7.08 |
1.621 |
09.39 |
11.55 |
09.13 |
2.418 |
09.26 |
12.31 |
9.51 |
2.727 |
14.11 |
1.69CaP_3xZrO2 |
09.62 |
7.14 |
1.873 |
08.51 |
08.96 |
07.02 |
1.556 |
06.61 |
10.57 |
8.11 |
2.266 |
10.85 |
1.69CaP_4xZrO2 |
11.62 |
8.23 |
2.853 |
11.04 |
09.76 |
07.46 |
1.989 |
09.97 |
09.62 |
7.15 |
1.993 |
10.89 |
1.69CaP_5xZrO2 |
11.46 |
8.06 |
2.897 |
09.70 |
08.10 |
06.13 |
1.435 |
05.43 |
08.53 |
6.44 |
1.578 |
07.06 |
Table 3 Structural parameters of HAP–ZrO2 composite powders determined from Rietveld refinement analysis after heat treatment at three different temperatures
Sample Code |
Structural parameters at 900 °C |
Structural parameters at 1000 °C |
Structural parameters at 1100 °C |
HAP |
t-ZrO2 |
HAP |
t-ZrO2 |
HAP |
t-ZrO2 |
a-axis (Å) |
c-axis (Å) |
a-axis (Å) |
c-axis (Å) |
a-axis (Å) |
c-axis (Å) |
a-axis (Å) |
c-axis (Å) |
a-axis (Å) |
c-axis (Å) |
a-axis (Å) |
c-axis (Å) |
1.67CaP_1xZrO2 |
9.5517(6) |
6.8342(4) |
3.5978(3) |
5.1782(7) |
9.5333(6) |
6.8422(4) |
3.5994(3) |
5.1794(6) |
9.4942(4) |
6.8517(3) |
3.5964(4) |
5.1792(4) |
1.67CaP_2xZrO2 |
9.5603(7) |
6.8253(5) |
3.5984(2) |
5.1789(5) |
9.5512(6) |
6.8336(4) |
3.5983(2) |
5.1823(4) |
9.4843(6) |
6.8625(5) |
3.5971(2) |
5.1816(5) |
1.67CaP_3xZrO2 |
9.5556(8) |
6.8289(6) |
3.5994(3) |
5.1787(6) |
9.5441(7) |
6.8353(5) |
3.5997(2) |
5.1797(5) |
9.4842(4) |
6.8568(3) |
3.5957(1) |
5.1792(2) |
1.67CaP_4xZrO2 |
9.5733(7) |
6.8204(6) |
3.5984(2) |
5.1814(4) |
9.5539(6) |
6.8296(4) |
3.5989(2) |
5.1815(4) |
9.4867(4) |
6.8586(4) |
3.5961(1) |
5.1807(2) |
1.67CaP_5xZrO2 |
9.5792(8) |
6.8128(7) |
3.5977(2) |
5.1801(4) |
9.5337(6) |
6.8363(5) |
3.5972(2) |
5.1805(5) |
9.4862(5) |
6.8594(4) |
3.5958(1) |
5.1821(2) |
1.68CaP_1xZrO2 |
9.5509(6) |
6.8359(4) |
3.5996(3) |
5.1786(8) |
9.4995(4) |
6.8560(3) |
3.5998(2) |
5.1784(5) |
9.4902(4) |
6.8549(3) |
3.5977(2) |
5.1784(4) |
1.68CaP_2xZrO2 |
9.5663(8) |
6.8307(6) |
3.6029(3) |
5.1776(7) |
9.4829(7) |
6.8482(5) |
3.5959(3) |
5.1719(5) |
9.4766(6) |
6.8491(5) |
3.5942(2) |
5.1703(4) |
1.68CaP_3xZrO2 |
9.5733(8) |
6.8203(6) |
3.5984(3) |
5.1808(6) |
9.4895(6) |
6.8545(5) |
3.5979(2) |
5.1775(4) |
9.4841(4) |
6.8563(4) |
3.5960(1) |
5.1783(3) |
1.68CaP_4xZrO2 |
9.5624(9) |
6.8275(7) |
3.6002(3) |
5.1787(6) |
9.4879(5) |
6.8554(4) |
3.5975(2) |
5.1775(3) |
9.4795(5) |
6.8578(4) |
3.5971(1) |
5.1769(3) |
1.68CaP_5xZrO2 |
9.5548(1) |
6.8155(9) |
3.5946(3) |
5.1748(7) |
9.4764(5) |
6.8588(5) |
3.5958(2) |
5.1815(2) |
9.4865(5) |
6.8560(5) |
3.5954(1) |
5.1814(3) |
1.69CaP_1xZrO2 |
9.5708(7) |
6.8256(5) |
3.6041(6) |
5.1753(9) |
9.5244(4) |
6.8402(3) |
3.6002(2) |
5.1750(6) |
9.4957(3) |
6.8580(2) |
3.5967(1) |
5.1840(4) |
1.69CaP_2xZrO2 |
9.5650(7) |
6.8263(5) |
3.6006(3) |
5.1790(6) |
9.5114(6) |
6.8499(4) |
3.6015(2) |
5.1771(5) |
9.4955(5) |
6.8546(4) |
3.5990(2) |
5.1783(4) |
1.69CaP_3xZrO2 |
9.5762(9) |
6.8203(7) |
3.6001(3) |
5.1808(7) |
9.5138(5) |
6.8466(4) |
3.5970(2) |
5.1808(3) |
9.4929(5) |
6.8548(4) |
3.5970(1) |
5.1793(3) |
1.69CaP_4xZrO2 |
9.5671(9) |
6.8257(9) |
3.6008(4) |
5.1816(7) |
9.5001(6) |
6.8528(5) |
3.5982(2) |
5.1808(4) |
9.4910(5) |
6.8555(4) |
3.5967(1) |
5.1785(3) |
1.69CaP_5xZrO2 |
9.5773(8) |
6.8175(9) |
3.6005(4) |
5.1800(8) |
9.5150(5) |
6.8428(5) |
3.5966(2) |
5.1818(3) |
9.4953(5) |
6.8510(4) |
3.5955(1) |
5.1812(3) |
3.3 Quantitative phase content
Table 4 represents the individual phase compositions of the powder samples obtained from Rietveld refinement at three different heat treatment temperatures, which have been synthesised by maintaining different Ca/P ratios and varying the ZrOCl2 precursor concentrations. From Table 4, it is observed that HAP and t-ZrO2 have occurred as the major phases, as expected, in all 15 samples after heat treatment at three different temperatures. At the heat treatment temperature of 900 °C, the phase content of t-ZrO2 has been found to increase in line with the corresponding increase in the amount of added ZrOCl2 precursor. In other words, the phase content of HAP has been found in decreasing order with the simultaneous increase in the content of t-ZrO2. An interesting feature is that the increase in the Ca/P molar ratio from 1.67 to 1.69 has also resulted in the gradual enhancement of the HAP phase formed in the composite powders. The synthesis performed using a Ca/P molar ratio of 1.67 and 1.69 has resulted in the formation of a minor amount of additional phase in the form of β-TCP other than HAP and t-ZrO2at 900 °C. All the previously reported work on the formation of HAP–ZrO2 composite powders has used yttria as a precursor to stabilize the t-ZrO2 phase.40–43 The present study has seen the stabilization of t-ZrO2 without the addition of any stabilizer during the synthesis. It has been reported by many authors that other than yttria, there are many additives in the form of CaO, MgO, La2O3 and rare earth elements that could be used to stabilize the t-ZrO2.44–47 It is also a well-known fact that depending on the extent of the calcium deficiency, the formation of β-TCP along with the HAP phase could not be avoided during heat treatment.4–6 Thus, it could be stated from the present investigation that the excess Ca2+ ions present in the calcium phosphate suspension has played an active role in the stabilization of t-ZrO2, thus creating a calcium deficiency and thereby leading to the formation of β-TCP during heat treatment. It should be noted that the weight fractions of all the β-TCP formed were found to be less than 5 wt%. In the case of the Ca/P precursors which had a maintained molar ratio of 1.68 during the synthesis, no β-TCP phase has been detected at the heat treatment temperature of 900 °C. Much research has been reported on the effect of the Ca/P molar ratio on the thermal stability and degradation behaviour of stoichiometric HAP. Osaka et al.48 states that the HAP synthesized by maintaining a Ca/P molar ratio of 1.68 has been found to be thermally stable until 1300 °C without phase degradation. Some of the reports also state that synthesis using a Ca/P molar ratio of >1.70 results in the non-stoichiometry of the HAP formed, which then gets thermally degraded into β-TCP and CaO beyond 1300 °C.49 Generally, the Ca/P molar ratio required to synthesize stoichiometric HAP is considered to be 1.67 and the resulting HAP starts to degrade beyond 1200 °C. Thus a minor alteration in the Ca/P molar ratio during the synthesis of stoichiometric HAP has resulted in the varied level of its thermal stability. However, the addition of ZrO2 precursor during the synthesis have resulted in the earlier degradation of the HAP phase, thus yielding a minor amount of β-TCP phase (less than 5 wt%) at 900 °C for the Ca/P molar ratios of 1.67 and 1.69. The addition of a varied range of ZrO2 precursors has not affected the HAP–ZrO2 composite powders at 900 °C for the synthesis performed with a Ca/P molar ratio of 1.68.
Table 4 Quantitative phase compositions of HAP–ZrO2 composite powders determined from Rietveld refinement analysis after heat treatment at three different temperatures
Sample Code |
Phase composition in weight fractions at 900 °C |
Phase composition in weight fractions at 1000 °C |
Phase composition in weight fractions at 1100 °C |
HAP |
t-ZrO2 |
β-TCP |
HAP |
t-ZrO2 |
β-TCP |
m-ZrO2 |
HAP |
t-ZrO2 |
β-TCP |
m-ZrO2 |
1.67CaP_1xZrO2 |
73.00 |
22.87 |
04.13 |
62.77 |
21.77 |
13.66 |
01.80 |
60.25 |
18.68 |
19.36 |
01.71 |
1.67CaP_2xZrO2 |
58.17 |
36.91 |
04.92 |
48.84 |
34.70 |
14.82 |
01.64 |
52.71 |
21.10 |
20.19 |
06.00 |
1.67CaP_3xZrO2 |
53.47 |
44.39 |
02.14 |
44.81 |
40.53 |
08.26 |
06.40 |
38.96 |
25.65 |
16.29 |
19.10 |
1.67CaP_4xZrO2 |
50.29 |
48.18 |
01.53 |
37.19 |
45.61 |
08.70 |
08.50 |
34.97 |
26.96 |
12.11 |
25.96 |
1.67CaP_5xZrO2 |
44.14 |
54.60 |
01.26 |
32.90 |
51.20 |
06.90 |
09.00 |
26.12 |
29.53 |
13.25 |
31.10 |
1.68CaP_1xZrO2 |
75.81 |
24.19 |
— |
64.52 |
20.13 |
12.53 |
02.82 |
63.66 |
19.55 |
15.86 |
00.93 |
1.68CaP_2xZrO2 |
62.68 |
37.32 |
— |
50.84 |
35.87 |
10.56 |
02.73 |
53.40 |
25.22 |
15.73 |
05.65 |
1.68CaP_3xZrO2 |
58.99 |
41.01 |
— |
46.00 |
39.66 |
07.94 |
06.40 |
39.00 |
31.26 |
14.98 |
14.76 |
1.68CaP_4xZrO2 |
52.72 |
47.28 |
— |
38.15 |
42.55 |
09.13 |
10.17 |
35.49 |
33.59 |
13.81 |
17.11 |
1.68CaP_5xZrO2 |
48.64 |
51.36 |
— |
34.75 |
46.00 |
08.25 |
11.00 |
28.99 |
24.01 |
15.37 |
31.63 |
1.69CaP_1xZrO2 |
77.32 |
20.40 |
02.28 |
67.46 |
20.85 |
09.49 |
02.20 |
66.51 |
16.69 |
11.96 |
04.84 |
1.69CaP_2xZrO2 |
64.20 |
33.72 |
02.08 |
56.46 |
33.34 |
07.74 |
02.46 |
55.28 |
26.85 |
11.23 |
06.64 |
1.69CaP_3xZrO2 |
59.64 |
38.82 |
01.54 |
49.12 |
40.18 |
05.23 |
05.47 |
41.93 |
32.77 |
11.58 |
13.72 |
1.69CaP_4xZrO2 |
53.35 |
44.80 |
01.85 |
39.18 |
41.67 |
07.78 |
11.37 |
36.80 |
35.05 |
10.63 |
17.52 |
1.69CaP_5xZrO2 |
50.52 |
46.81 |
02.67 |
36.16 |
43.84 |
08.00 |
12.00 |
29.13 |
36.38 |
10.66 |
23.83 |
Heat treatment at 1000 °C and 1100 °C has revealed a significant change in the phase composition of all the powder samples. Despite the varied level of the Ca/P molar ratio maintained during the synthesis, the HAP–ZrO2 composite powders have shown phase degradation at both the temperatures of 1000 °C and 1100 °C by the formation of a considerable amount of m-ZrO2 (Table 4). The phase content of the β-TCP formed has been found to be comparably higher at 1000 °C and 1100 °C when compared to their content at 900 °C. The phase content of the newly formed m-ZrO2 phase at 1100 °C has been seen at higher levels when compared to its content at 1000 °C, thus signifying that there is a phase degradation of HAP–ZrO2 composite powders at 1000 °C. Furthermore, with the increased addition of ZrO2 precursor during synthesis, the phase content of m-ZrO2 has been found to increase. It is important to mention here that in any of the compositions investigated at three different heat treatment temperatures, neither c-ZrO2 nor the formation of CaO has been detected in trace levels. The results obtained from the present investigation are a significant contradiction to the reported results shown in eqn (1) above, which shows that the degradation of HAP–ZrO2 composites is initiated by the formation of the m-ZrO2 and CaO phases after the heat treatment temperature of 950 °C. Reports in the literature have proposed a different kind of mechanism for the formation of m-ZrO2 and c-ZrO2 during the degradation of HAP–ZrO2 composites. The formation of c-ZrO2 has been attributed to the decomposition of HAP to yield CaO, which in turn is consumed by the t-ZrO2 to be converted into c-ZrO2.50,51 The formation of m-ZrO2 has been attributed to the presence of an insufficient amount of CaO from the degradation of HAP to stabilize the c-ZrO2.52 The comparison of the phase content of the different compositions in Table 4 indicates that the enhanced content of m-ZrO2 at 1100 °C causes the rapid degradation of the composite powders when compared with the degradation behaviour at 1000 °C. A greater amount of the phase content of m-ZrO2 has been formed with the corresponding increase in the ZrO2 content in the samples. Thus, the results from the present investigation infers that the ZrO2 content has an active role in the degradation behaviour of HAP–ZrO2 composite powders.
3.4 Raman spectroscopy
Fig. 4–6 represent the Raman spectra of the HAP–ZrO2 composite powders after heat treatment at different temperatures, together with the spectra of pure HAP. From Fig. 4–6, it is observed that the bands representing the PO4 groups of the HAP phase53,54 have also been seen in pure HAP and also for all the HAP–ZrO2 composite powders. The bands observed around the region of 426 cm−1 and 458 cm−1 correspond to the ν2 bending vibrations of the PO43− ion. The bands observed in the region of 578, 589 and 610 cm−1 correspond with the ν4 fundamental vibrational modes that arise from the triply degenerate bending vibrations. The high intensity band seen at 958 cm−1 has arisen from the symmetric stretching of the ν1 fundamental vibrational modes. The bands located at 1033, 1049 and 1073 cm−1 have been attributed to the asymmetric stretching vibrations of the ν3 fundamental vibrational modes of the PO4 bonds. The Raman spectra for the powders after heat treatment at 900 °C (Fig. 4) confirm the bands corresponding to the t-ZrO2 phase.
 |
| Fig. 4 Raman spectra for the HAP–ZrO2 composite powders heat treated at 900 °C with the synthesis done at a Ca/P molar ratio of 1.68. | |
The bands observed in the region around 266, 146 and 312 cm−1 correspond with the presence of t-ZrO2.55,56 At the heat treatment temperature of 900 °C, the intensity of the characteristic bands of the HAP phase has decreased with the corresponding increase in the intensity of bands of the t-ZrO2 phase.
For Raman spectra obtained for the powder samples after heat treatment at 1000 °C and 1100 °C (Fig. 5 and 6), the bands corresponding to the HAP and t-ZrO2 phases have been identified in the spectrum. Other than the bands of the HAP and t-ZrO2 phases, the bands corresponding to the presence of m-ZrO2 phase have been confirmed from the bands witnessed in the spectral region around 178, 189, 333, 347 and 380 cm−1.57,58 Since the overall quantitative yield of m-ZrO2 has been found to a lesser extent at 1000 °C when compared with its content at 1100 °C, the bands of m-ZrO2 at 1000 °C have been seen with less intensity towards 1100 °C. Additionally, the bands of m-ZrO2 have been observed with increasing intensity, with their increasing content at 1100 °C. Thus, the results obtained from the Raman spectroscopy have been found to be in agreement with the quantitative analysis from the Rietveld refinement.
 |
| Fig. 5 Raman spectra for the HAP–ZrO2 composite powders heat treated at 1000 °C with the synthesis done at a Ca/P molar ratio of 1.68. | |
 |
| Fig. 6 Raman spectra for the HAP–ZrO2 composite powders heat treated at 1100 °C with the synthesis done at a Ca/P molar ratio of 1.68. | |
Summary
A wide range of HAP–ZrO2 composite powders have been obtained in the present investigation by adapting the in situ aqueous precipitation technique. Minor alterations in the Ca/P molar ratio together with five different precursor concentrations of ZrO2 have been varied during the synthesis. Heat treatment of the resulting powders has induced the formation of an HAP–ZrO2 composite and the resulting phase content has been found to vary depending on the molar ratio of the Ca/P precursors and the concentration of the ZrOCl2 precursor. A minor amount of β-TCP (less than 5 wt%) together with the HAP and t-ZrO2 has been detected for the Ca/P precursors with molar ratios of 1.67 and 1.69. For the precursor with the Ca/P molar ratio of 1.68, however, the formation of HAP and t-ZrO2 was only confirmed after heat treatment at 900 °C. Heat treatment at 1000 °C and 1100 °C has confirmed the phase degradation of the HAP–ZrO2 composite powders by virtue of the formation of significant amounts of m-ZrO2. The results from the present study show that the stabilization of the t-ZrO2 phase had occurred without the addition of any stabilizing agent in the form of yttria, which has been a usual practice during the formation of HAP–ZrO2 composites. In the absence of any stabilizing agent, the presence of Ca2+ ions in the calcium phosphate suspension had played a significant role in the t-ZrO2 phase stabilization during heat treatment. The increased addition of ZrOCl2 precursor had resulted in the corresponding enhancement in the phase content of ZrO2 in the composites with the simultaneous reduction in the HAP content. The presence of a higher content of ZrO2 in the composites had led to the enhanced formation of m-ZrO2, thus signifying the rapid degradation of the composite powders.
Acknowledgements
The authors gratefully acknowledge the financial support received from the Council of Scientific and Industrial Research, India (CSIR Scheme no. 22(612)/12/EMR-II). The authors are also grateful for the use of the instrumentation at the Central Instrumentation Facility (CIF) of Pondicherry University.
References
- L. M. Rodriguez-Lorenzo, J. N. Hart and K. A. Gross, Structural and chemical analysis of well-crystallized hydroxyfluorapatites, J. Phys. Chem. B, 2003, 107, 8316–8320 CrossRef CAS.
- J. C. Elliot, Structure and chemistry of the apatites and other calcium orthophosphates, Elsevier, Amsterdam, 1994 Search PubMed.
- M. Jarcho, C. H. Bolen, M. B. Thomas, J. Bobick, J. F. Kay and R. H. Doremus, Hydroxyapatite synthesis and characterization in dense polycrystalline form, J. Mater. Sci., 1976, 11, 2027–2035 CrossRef CAS.
- S. Kannan, S. I. Vieira, S. M. Olhero, P. M. C. Torres, S. Pina, O. A. B. da Cruz e Silva and J. M. F. Ferreira, Synthesis, mechanical and biological characterization of ionic doped carbonated hydroxyapatite/β-tricalcium phosphate mixtures, Acta Biomater., 2011, 7, 1835–1843 CrossRef CAS PubMed.
- S. Kannan, A. F. Lemos and J. M. F. Ferreira, Synthesis and mechanical performance of biological-like hydroxyapatites, Chem. Mater., 2006, 18, 2181–2186 CrossRef CAS.
- S. Kannan and J. M. F. Ferreira, Synthesis and thermal stability of hydroxyapatite-β-tricalcium phosphate composites with cosubstituted sodium, magnesium, and fluorine, Chem. Mater., 2006, 18, 198–203 CrossRef CAS.
- G. Willmann, Coating of Implants with Hydroxyapatite Material Connections between Bone and Metal, Adv. Eng. Mater., 1999, 1, 95–105 CrossRef CAS.
- J. L. Ong and D. C. N. Chan, Hydroxyapatite and their use as coatings in dental implants, A Review, Critical Reviews™ in Biomedical Engineering, 1999, vol. 28, pp. 1–41 Search PubMed.
- L. L. Hench and E. C. Ethridge, Biomaterials – An Interfacial Approach, Academic Press, New York, 1982 Search PubMed.
- M. D. Rajiv Gandhi, J. M. D. Roderick Davey, N. Nizar and M. D. Mahomed, Hydroxyapatite Coated Femoral Stems in Primary Total Hip Arthroplasty, J. Arthroplasty, 2009, 24, 38–42 CrossRef PubMed.
- B. D. Ratner, A. S. Hoffman, F. J. Schoen and J. E. Lemons, Biomaterials Science, An Introduction to Materials in Medicine, Academic press, 2004 Search PubMed.
- Z. Xihua, L. Changxia, L. Musen, B. Yunqiang and S. Junlong, Fabrication of hydroxyapatite/diopside/alumina composites by hot-press sintering process, Ceram. Interfaces, 2009, 35, 1969–1973 CrossRef PubMed.
- E. Adolfsson, M. Nygren and L. Hermansson, Decomposition mechanisms in aluminium oxide–apatite systems, J. Am. Ceram. Soc., 1999, 82, 2909–2912 CrossRef CAS PubMed.
- H. Kurzweg, R. B. Heimann, M. Troczynski and M. L. Wayman, Development of plasma-sprayed bioceramic coatings with bond coats based on titania and zirconia, Biomaterials, 1998, 19, 1507–1511 CrossRef CAS.
- H.-W. Kim, H.-E. Kim, V. Salih and J. C. Knowles, Hydroxyapatite and titania sol–gel composite coatings on titanium for hard tissue implants; mechanical and in vitro biological performance, J. Biomed. Mater. Res., Part B, 2005, 72, 1–8 CrossRef PubMed.
- S. Nath, R. Tripathi and B. Basu, Understanding phase stability, microstructure development and biocompatibility in calcium phosphate–titania composites, synthesized from hydroxyapatite and titanium powder matrix, Mater. Sci. Eng., C, 2009, 29, 97–107 CrossRef CAS PubMed.
- C. Piconi and G. Maccauro, Zirconia as a ceramic biomaterial-review, Biomaterials, 1999, 20, 1–25 CrossRef CAS.
- S. Affatato, M. Goldoni, M. Testoni and A. Toni, Mixed oxides prosthetic ceramic ball heads. Part 3: effect of the ZrO2 fraction on the wear of ceramic on ceramic hip joint prostheses. A long-term in vitro wear study, Biomaterials, 2001, 22, 717–723 CrossRef CAS.
- B. Basu, J. Vleugels and O. Van Der Biest, Transformation behaviour of tetragonal zirconia: role of dopant content and distribution, Mater. Sci. Eng., A, 2004, 366, 338–347 CrossRef PubMed.
- D. Singh, M. de la Cinta Lorenzo-Martin, F. Gutierrez-Mora, J. L. Routbort and E. D. Case, Self-joining of zirconia–hydroxyapatite composites using plastic deformation
process, Acta Biomater., 2006, 2, 669–675 CrossRef CAS PubMed.
- H.-W. Kim, Y.-H. Koh, B.-H. Yoon and H.-E. Kim, Reaction sintering and mechanical properties of hydroxyapatite–zirconia composites with calcium fluoride additions, J. Am. Ceram. Soc., 2002, 85, 1634–1636 CrossRef CAS PubMed.
- Z. E. Erkmen, Y. Gen and F. N. Oktar, Microstructural and mechanical properties of hydroxyapatite-zirconia composites, J. Am. Ceram. Soc., 2007, 90, 2885–2892 CrossRef CAS PubMed.
- E. S. Ahn, N. J. Gleason and J. Y. Ying, The effect of zirconia reinforcing agents on the microstructure and mechanical properties of hydroxyapatite-based nanocomposites, J. Am. Ceram. Soc., 2005, 88, 3374–3379 CrossRef CAS PubMed.
- J. Zhang, M. Iwasa, N. Kotobuki, T. Tanaka, M. Hirose, H. Ohgushi and D. Jiang, Fabrication of hydroxyapatite–zirconia composites for orthopedic applications, J. Am. Ceram. Soc., 2006, 89, 3348–3355 CrossRef CAS PubMed.
- A. Rapacz-Kmita, A. Slusarczyk, Z. Paszkiewicz and C. Paluszkiewicz, Phase stability of hydroxyapatite–zirconia (HAP–ZrO2) composites for bone replacement, J. Mol. Struct., 2004, 704, 333–340 CrossRef CAS PubMed.
- E. Adolfsson, P. Alberius-Henning and L. Hermansson, Phase analysis and thermal stability of hot isostatically pressed zirconia–hydroxyapatite composites, J. Am. Ceram. Soc., 2000, 83, 2798–2802 CrossRef CAS PubMed.
- W. Li and L. Gao, Fabrication of HAp–ZrO2 (3Y) nano-composite by SPS, Biomaterials, 2003, 24, 937–940 CrossRef CAS.
- Z. Shen, E. Adolfsson, M. Nygren, L. Gao, H. Kawaoka and K. Niihara, Dense hydroxyapatite–zirconia ceramic composites with high strength for biological applications, Adv. Mater., 2001, 13, 214–216 CrossRef CAS.
- A. C. Larson and R. B. Von Dreele, General Structure Analysis System (GSAS), Los Alamos National Laboratory Report LAUR, 2004, pp. 86–748 Search PubMed.
- B. H. Toby, EXPGUI: A Graphical User Interface for GSAS, J. Appl. Crystallogr., 2001, 34, 210–213 CrossRef CAS.
- R. M. Wilson, J. C. Elliot and S. E. P. Dowker, Rietveld refinement of the crystallographic structure of human dental enamel apatites, Am. Mineral., 1999, 84, 1406–1414 CAS.
- C. J. Howard, R. J. Hill and B. E. Reichert, Structures of the ZrO2 polymorphs at room temperature by high-resolution neutron powder diffraction, Acta Crystallogr., Sect. B: Struct. Sci., 1988, 44, 116–120 CrossRef.
- M. Yashima, S. Takashi, A. Kamiyama and A. Hoshikawa, Crystal structure analysis of β-tricalcium phosphate Ca3(PO4)2 by neutron powder diffraction, J. Solid State Chem., 2003, 175, 272–277 CrossRef CAS.
- D. K. Smith and H. W. Newkirk, The crystal structure of baddeleyite (monoclinic ZrO2) and its relation to the polymorphism of ZrO2, Acta Crystallogr., 1965, 18, 983–991 CrossRef CAS.
- S. Vasanthavel, P. Nandha Kumar and S. Kannan, Quantitative analysis on the influence of SiO2 content on the phase behavior of ZrO2, J. Am. Ceram. Soc., 2014, 97, 635–642 CrossRef CAS PubMed.
- E. S. Ahn, N. J. Gleason, A. Nakahira and J. Y. Ying, Nanostructure processing of hydroxyapatite-based bioceramics, Nano Lett., 2001, 1, 149–153 CrossRef CAS.
- H. Guoa, K. A. Khorb, Y. C. Boeya and X. Miaoa, Laminated and functionally graded hydroxyapatite/yttria stabilized tetragonal zirconia composites fabricated by spark plasma sintering, Biomaterials, 2003, 24, 667–675 CrossRef.
- P. Nandha Kumar, M. Boovarasan, R. Kishore Singh and S. Kannan, Synthesis, structural analysis and fabrication of coatings of the Cu2+ and Sr2+ co-substitutions in β-Ca3(PO4)2, RSC adv., 2013, 3, 22469–22479 RSC.
- E. Wu, E. H. Kisi and E. A. Gray, Modelling dislocation-induced anisotropic line broadening in rietveld refinements using a voigt function. II. Application to neutron powder diffraction data, Appl. Crystallogr., 1998, 31, 363–368 CrossRef CAS.
- Y.-M. Sung, Y.-K. Shin and J.-J. Ryu, Preparation of hydroxyapatite–zirconia bioceramic nanocomposites for orthopaedic and dental prosthesis applications, Nanotechnology, 2007, 18, 065602 CrossRef.
- Z. Evis, C. Ergun and R. H. Doremus, Hydroxylapatite–zirconia composites: thermal stability of phases and sinterability as related to the CaO–ZrO2 phase diagram, J. Mater. Sci., 2005, 40, 1127–1134 CrossRef CAS.
- M. R. Towler, I. R. Gibson and S. M. Best, Novel processing of hydroxyapatite–zirconia composites using nano-sized particles, J. Mater. Sci. Lett., 2000, 19, 2209–2211 CrossRef CAS.
- Z. Evis, M. Sato and T. J. Webster, Increased osteoblast adhesion on nanograined hydroxyapatite and partially stabilized zirconia composites, J. Biomed. Mater. Res., Part A, 2006, 78, 500–507 CrossRef PubMed.
- R. C. Garvie and P. S. Nicholson, Structure and thermodynamical properties of partially stabilized zirconia in the CaO–ZrO2 system, J. Am. Ceram. Soc., 1972, 55, 152–157 CrossRef CAS PubMed.
- R. C. Garvie, C. Urbani, D. R. Kennedy and J. C. Mcneuer, Biocompatibility of magnesia partially stabilized zirconia (Mg-PSZ) ceramics, J. Mater. Sci., 1984, 19, 3224–3228 CrossRef CAS.
- P. Thangadurai, A. Chandra Bose and S. Ramasamy, Phase stabilization and structural studies of nanocrystalline La2O3–ZrO2, J. Mater. Sci., 2005, 40, 3963–3968 CrossRef CAS PubMed.
- I. Denrya and J. Robert Kelly, State of the art of zirconia for dental applications, Dent. Mater., 2008, 24, 299–307 CrossRef PubMed.
- A. Osaka, Y. Miura, K. Takeuchi, M. Asada and K. Takahashi, Calcium apatite prepared from calcium hydroxide and orthophosphoric acid, J. Mater. Sci.: Mater. Med., 1991, 2, 51–55 CrossRef CAS.
- S. Raynaud, E. Champion, D. Bernache-Assollant and P. Thomas, Calcium phosphate apatites with variable Ca/P atomic ratio I. Synthesis, characterization and thermal stability of powders, Biomater., 2002, 23, 1065–1072 CrossRef CAS.
- K. Ioku, S. Somiya and M. Yoshimura, Hydroxyapatite Ceramics with Tetragonal Zirconia Particles Dispersion Prepared by HIP Post-Sintering, J. Ceram. Soc. Jpn Int. Ed., 1991, 99, 196–203 CrossRef CAS.
- K. Yamashita, T. Kobayashi, M. Kitamura, T. Umegaki and T. Kanazawa, Effect of Water Vapor on the Solid-State Reaction between Hydroxyapatite and Zirconia or CaO-PSZ, J. Ceram. Soc. Jpn., 1988, 96, 616–619 CrossRef CAS.
- M. Inuzuka, S. Nakamura, S. Kishia, K. Yoshida, K. Hashimoto, Y. Toda and K. Yamashita, Hydroxyapatite-doped zirconia for preparation of biomedical composites ceramics, Solid State Ionics, 2004, 172, 509–513 CrossRef CAS PubMed.
- A. Jillavenkatesa and R. A. Condrate Sr., Sol–gel processing of hydroxyapatite, J. Mater. Sci., 1998, 33, 4111–4119 CrossRef CAS.
- S. Koutsopoulos, Synthesis and characterization of hydroxyapatite crystals: a review study on the analytical methods, J. Biomed. Mater. Res., 2002, 62, 600–612 CrossRef CAS PubMed.
- B. Yılmaz and Z. Evis, Raman spectroscopy investigation of nano hydroxyapatite doped with yttrium and fluoride ions, Spectrosc. Lett., 2014, 47, 24–29 CrossRef.
- D.-J. Kim, H.-J. Jung and I.-S. Yang, Raman spectroscopy of tetragonal zirconia solid solutions, J. Am. Ceram. Soc., 1993, 76, 2106–2109 CrossRef CAS PubMed.
- E. F. Lopez, V. S. Escribano, M. Panizza, M. M. Carnasciali and G. Busca, Vibrational and electronic spectroscopic properties of zirconia Powders, J. Mater. Chem., 2001, 11, 1891–1897 RSC.
- C. Carlone, Raman spectrum of Zirconia–Hafnia mixed crystals, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 2079–2084 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2014 |
Click here to see how this site uses Cookies. View our privacy policy here.