You are on page 1of 8

Particuology 9 (2011) 537544

Contents lists available at SciVerse ScienceDirect

Particuology
journal homepage: www.elsevier.com/locate/partic

Characterization of single-crystal uorapatite nanoparticles synthesized via mechanochemical method


Reza Ebrahimi-Kahrizsangi, Bahman Nasiri-Tabrizi , Akbar Chami
Materials Engineering Department, Najafabad Branch, Islamic Azad University, Najafabad, Isfahan, Iran

a r t i c l e

i n f o

a b s t r a c t
The synthesis of nanostructured uorapatite (FA; Ca10 (PO4 )6 F2 ) was explored from the starting materials of CaHPO4 , Ca(OH)2 , CaO, P2 O5 and CaF2 via a mechanochemical process. In this research, the suitability of using the mechanochemical process to prepare a high crystalline phase of FA was studied. The characterization and structural features of the synthesized powders were evaluated using powder X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FT-IR), energy dispersive X-ray spectroscopy (EDX), scanning electron microscopy (SEM), and transmission electron microscopy (TEM) techniques. The results from the structural studies indicate that the maximum lattice disturbance in the apatite structure after the mechanochemical process was at the (0 0 2) plane. Furthermore, the maximum particle size was below the crystallite size after 60 h of milling and subsequent thermal treatment at 600 C for 1 h (heated up to 600 C and kept for 1 h at this temperature). We determined that this method gives rise to the single-crystal FA with an average size in the range of 25 5 to 29 9 nm. The present ndings suggest that the solid-state reaction and appropriate thermal process simultaneously lead to the formation of nanostructured FA with spheroidal shape. 2011 Chinese Society of Particuology and Institute of Process Engineering, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved.

Article history: Received 3 June 2011 Received in revised form 7 July 2011 Accepted 27 July 2011 Keywords: Fluorapatite Single-crystal Nanoparticles Mechanochemical Crystallinity

1. Introduction Hydroxyapatite (HA, Ca10 (PO4 )6 (OH)2 ) has been used for biomedical applications because of its superior biocompatibility and bioactivity (Roeder, Converse, Leng, & Yue, 2006). Among the different forms of calcium phosphates, particular attention has been placed on HA because the bone mineral consists of tiny HA crystals (Kalita, Bhardwaj, & Bhatt, 2007). Nevertheless, HA has a high dissolution rate in biological systems (Fini et al., 2003), poor corrosion resistance in an acid environment and poor chemical stability at high temperature (Chen & Miao, 2005), which has restricted wider applications in the elds of orthopedics and dentistry. The inorganic matrix of the bone is based on HA doped with different quantities of cations, such as Na+ , K+ and Mg2+ , and anions, such as CO3 2 , SO4 2 and F . Among them, F plays a leading role because of its inuence on the physical and biological characteristics of HA (Nikcevic et al., 2004). When uoride (F) is consumed in optimal amounts in water and food and used topically in toothpaste, mouth rinses, and professionally applied ofce treatments, it increases tooth mineralization and bone density,

Corresponding author. Tel.: +98 3114456551; fax: +98 3312291008. E-mail address: bahman nasiri@hotmail.com (B. Nasiri-Tabrizi).

reduces the risk and prevalence of dental caries and helps to promote enamel remineralization throughout life for individuals of all ages (Palmer & Anderson, 2001). Therefore, uorine-substituted HA (FHA, Ca10 (PO4 )6 (OH)2x Fx ) has attracted substantial attention as a clinical restoration material for improvement of the biostability of HA in biomedical applications (Fini et al., 2003). This substitution causes an increase in chemical stability, a decrease in mineral solubility, and a promotion in bone cell proliferation (Barinov, Shvorneva, Ferro, Fadeeva, & Tumanov, 2004; Jantova, Theiszova, Letasiova, Birosova, & Palou, 2008). If the OH in HA is completely substituted by F , uorapatite (FA, Ca10 (PO4 )6 F2 ) is formed. FA is found in dental enamel and is usually used in dental applications due to its greater mechanical strength. For all these reasons, the synthesis of FHA and FA is of great value and has been widely investigated by multiple techniques, such as precipitation (Chen & Miao, 2005), solgel (Cheng, Zhang, & Weng, 2006), hydrolysis (Kurmaev et al., 2002), hydrothermal (Rodriguez-Lorenzo, Hart, & Gross, 2003), and mechanochemical methods (Zhang, Zhu, & Xie, 2005). Over the past decades, the mechanosynthesis process has been developed for the production of a wide range of nanostructured materials (Suryanarayana, 2001). The advantages of this technique are that melting is not necessary and that the powders are nanocrystalline (Silva, Graca, Valente, & Sombra, 2007). In general, the mechanochemical process can be performed under wet or dry conditions (Rhee, 2002; Silva, Pinheiro, Miranda, Goes, &

1674-2001/$ see front matter 2011 Chinese Society of Particuology and Institute of Process Engineering, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved.

doi:10.1016/j.partic.2011.07.001

538

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

Sombra, 2003). In most cases, the wet routes require precise controls on the processing conditions, and the characteristics of the nal products are greatly inuenced by various parameters. In addition, the synthesis procedure and apparatus are complicated. These difculties lead to the synthesis of uoridated HA with poor reproducibility and high processing costs, so it is not suitable for mass production (Rhee, 2002; Zhang et al., 2005). However, in the dry mechanosynthesis process, many of the difculties are eliminated because the reactions are faster without the addition of water and a very low level of pollution by the mill material is observed (El Briak-BenAbdeslam, Ginebra, Vert, & Boudeville, 2008). Moreover, this process has the benet of high reproducibility and low processing costs (Rhee, 2002). In most papers and patents concerning the synthesis of uoridated HA, the mechanochemical process was performed (Fathi & Mohammadi Zahrani, 2009a, 2009b; Mohammadi Zahrani & Fathi, 2009; Silva et al., 2003), and only a few research groups have been devoted to the mechanochemical process and subsequent thermal treatment on FA bioceramic (Barinov et al., 2004). The present work aims to synthesize nanocrystalline FA (Ca10 (PO4 )6 F2 ) using novel solid-state processes that consist of the dry mechanochemical technique (to obtain nanostructured FA using high purity powders) and the thermal annealing process (for recovery of crystallinity degree of apatite powders). It was found that the control of the quality and crystallinity of apatite phases are important for improvement of their dissolution in the human body (Nakano, Tokumura, & Umakoshi, 2002). Therefore, studying the crystallinity, morphological characteristics and other structural features (crystallite size, lattice strain) of FA are important to understand its suitability for different biomedical applications. 2. Materials and methods 2.1. Preparation method FA nanopowders with different structural characteristics were synthesized through novel dry mechanochemical processes, which are summarized in Fig. 1. The purpose of the milling was twofold: rst, to activate the following reactions via mechanochemical processes, and second, to produce the nanostructured FA.

The starting reactants were anhydrous dicalcium phosphate (CaHPO4 , Merck, Darmstadt, Germany), calcium hydroxide (Ca(OH)2 , SigmaAldrich Corporation, St. Louis, USA), phosphorous pentoxide (P2 O5 , Merck, Darmstadt, Germany), calcium oxide (CaO, Merck, Darmstadt, Germany) and calcium uoride (CaF2 , Merck, Darmstadt, Germany). To synthesize the nanostructured FA, the initial powders were ground on a high energy planetary mill with the stoichiometric proportionality between the materials given in following reactions: 6CaHPO4 + 3Ca(OH)2 + CaF2 Ca10 (PO4 )6 F2 + 6H2 O 9CaO + 3P2 O5 + CaF2 Ca10 (PO4 )6 F2 (I) (II)

Milling processes were performed in polyamide 6 vials (150 mL) using zirconia balls (diameter 20 mm) under ambient air atmosphere. The charge-to-ball ratio and rotational speed were 1:20 and 600 rpm, respectively. To prevent excessive heating, the milling was carried out in 45-min milling steps with 15-min interval pauses. The subsequent thermal annealing process was performed in an electrical furnace at 600 C for 1 h in air. 2.2. Characterization Phase analysis and structural transformation were evaluated with the Philips X-ray diffractometer (XRD) with Cu K radiation. The diffractometer was operated at 40 kV and 30 mA. All measurements were performed at room temperature, 25 C, within a diffraction range of 2 = 060 at 1 /s speed. PANalytical XPert HighScore software (Almelo, The Netherland) was also used for the analysis of different peaks. The diffraction patterns of products were compared to standards proposed by the Joint Committee on Powder Diffraction and Standards (JCPDS), which involved card # 15-0876 for FA, # 24-0033 for HA, and # 09-0080 for CaHPO4 . The functional groups of ne powders were examined using Fourier transform infrared spectroscopy (FT-IR: Bruker, TENSOR27). One milligram of the powdered sample was mixed with 200 mg spectroscopic grade KBr by hand-milling the powder in an agate mortar. Subsequently, the transmittance spectrum was recorded in the range 4000400 cm1 at 2 cm1 resolution by 16 scans. Scanning electron microscopy (SEM, SERON AIS-2100) was applied to characterize the morphological characteristics of the samples, which were sputter-coated with a thin layer of gold with a PVD apparatus (Fathi & Hani, 2007). The accelerating voltage and vacuum control were 20 kV and fully automated, respectively. Energy dispersive X-ray spectroscopy (EDX), which was coupled with SEM, was used for semi-quantitative examination of the samples (voltage used for EDX equal to 20 kV). A dilute suspension of the powders obtained in ethanol were collected on carbon-coated copper grids and allowed to dry for 3 min. The size and morphology of ne powders were observed in a transmission electron microscope (Philips CM10, Eindhoven, The Netherlands) operated at 100 kV. 3. Results and discussion 3.1. XRD analyses In Fig. 2(a) and (b), XRD patterns of the FA1 and FA2 samples are shown, respectively, showing that the powders synthesized through the two different mechanochemical processes are mostly FA. The JCPDS reference 15-0876 was used, and major peaks between 2 = 2060 were obtained. The peaks at 2 = 25.86, 31.94, 32.27, 33.13, 34.14, 40.04, 46.87, and 49.58 were compared and

Fig. 1. Flow sheet of FA nanoparticles preparation.

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

539

contrasted to FA1 shown in Fig. 2(a) and FA2 (Fig. 2(b)). Complete agreement with the standard card of uorapatite (JCPDS #15-0876) was not observed for FA1 due to the presence of additional peaks at 2 = 26.59 and 30.19 . These additional peaks were attributed to CaHPO4 from the starting materials. In FA2, complete agreement with JCPDS #15-0876 was observed, allowing FA2 to be used as the pure FA phase sample when required. On the other hand, in sample FA1, diffraction peaks were broadened, especially between 2 = 3134 , indicating that the synthesized sample exhibited poor crystallinity. However, FA2 exhibited a more ordered structure. The results indicate that the degree of crystallinity of the single phase FA obtained through mechanosynthesis depends on the initial synthesis conditions and chemical composition (e.g., deformations in the structure of FA1 due to Ca-lead amorphousity, which was not observed in the undeformed FA2 sample). In the unsubstituted parent compound (HA), the decomposition prole is anticipated to be different from the substituted derivatives (FA or FHA). For HA, when heated at a rate of 5 C/min in the temperature range of 600800 C, decomposition occurs. In general, the decomposition temperature strongly depends on the characteristics and synthetic technique of the apatite powder (Fathi & Hani, 2007). To obtain pure FA with a highly crystalline structure, the annealing temperature was selected to be 600 C for 1 h. In Fig. 2(c) and (d), XRD patterns of the FA3 and FA4 samples after the heat treatment are shown, respectively. According to Fig. 2(c), broadened crystallographic reection peaks at 2 = 32.27 and 33.13 appeared after milling and changed into sharp peaks after annealing. Moreover, the diffraction peaks of the extra phase (CaHPO4 ) diminished after the annealing process. A similar thermal recovery process of crystallinity was observed in sample FA4, and a highly crystalline structure of apatite was obtained (Fig. 2(d)). It should be noted that the crystallinity degree depends not only on the heat treatment conditions but also on the amount of strain introduced during the milling process. Therefore, the evaluation of crystallite size, lattice strain, and lattice parameters to determine the crystallinity degree is necessary (De Castro & Mitchell, 2002). Based on the available crystallographic data, FA has a very similar atomic structure and belongs to the same space and group as HA (space group: P63 /m; parameters: a = b = 9.462 A = = 90 , = 120 ), differing only in the substitution c = 6.849 A, of F for the OH groups in the HA structure. Because F is smaller than OH , the substitution results in a contraction in the a-axis dimensions to 0.9368 nm but with no change in the c-axis dimension. The d-spacing and diffraction peak intensity of the prepared powders were compared with the JCPDS standard corresponding to FA (JCPDS 15-0876) (Table 1). There is a good consistency between the calculated and standard values. Moreover, Table 1 shows that the (h k l = 2 1 1) peak intensity of FA maintained a given level (100%) after heat treatment at 600 C, which indicates that the FA has a stable structure at 600 C and conrms the formation of a stoichiometric phase. The crystallite size and lattice strain of the powder samples were determined with the XRD data according to the Scherrer formula and the tangent formula (Suryanarayana, 2001):

Fig. 2. XRD patterns of the samples milled for 60 h in polymeric vials: (a) FA1, (b) FA2, (c) FA3 and (d) FA4.

D=

K , B cos

(1)

E=

B . 4 tan

(2)

540

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

Table 1 d-Spacing of the prepared samples in comparison with the standard FA (JCPDS #15-0876). d-Spacing (nm) FA1 0.34452 0.28190 0.27168 0.19426 0.18503 0.17269 FA2 0.34391 0.28035 0.27059 0.19359 0.18367 0.17164 FA3 0.34449 0.28062 0.27015 0.19350 0.18384 0.17221 FA4 0.34417 0.28038 0.27069 0.19377 0.18366 0.17209 JCPDS 0.34420 0.28000 0.27020 0.19370 0.18370 0.17220 Relative intensity (%) FA1 28.4 100 77.1 54.4 47.45 39.42 FA2 29.7 100 59.6 29 42.2 18.1 FA3 28.9 100 56.5 42.1 50 24.84 FA4 31.6 100 65.9 34.9 45.8 19.9 JCPDS 40 100 60 25 30 16 (0 0 2) (2 1 1) (3 0 0) (2 2 2) (2 1 3) (0 0 4) hkl

Table 2 Average crystallite size, lattice strain and crystallinity of the samples. Sample FA1 FA2 FA3 FA4 Average crystallite size, D (nm) 25 35 36 36 6 3 2 3 Average lattice strain, E (%) 0.434 0.304 0.333 0.291 Crystallinity, Xc (%) 45 71 89 93

B describes the structural broadening, which is the difference in the integral prole widths between a standard and the observed sample: B(size) = Bobs. Bstd. ,
2 2 B2 (strain) = Bobs. Bstd. ,

On the other hand, the relationship between lattice spacing (d) and lattice parameters (a, b, and c) of the hexagonal structure is 1 4 h2 + hk + k2 l2 = + 2, 3 d2 a2 c (6)

(3) (4)

where , D, E, and are the wavelength of the X-ray used (0.154056 nm), crystallite size (average), lattice strain (mean lattice distortion) and Bragg angle ( ), respectively. The average values of the crystallite size and lattice strain were determined for the (2 1 1), (3 0 0), (2 1 3) and (2 2 2) Millers planes. In Table 2, the values of average crystallite size and lattice strain determined for both samples indicate that the thermal annealing process at approximately 600 C is not accompanied by a remarkable change in crystallite size or lattice strain. The relative crystallization percentage is measured using a method developed by Landi, Tampieri, Celotti, and Sprio (2000). This method compares the intensity of the FA peak at (3 0 0) plane (I300 ) with the intensity of the hollow between peaks of (1 1 2) and (3 0 0) planes (V1 1 2/3 0 0 ): V1 1 2/3 0 0 I3 0 0

Xc = 1

(5)

As shown in Table 2, the fraction of crystalline phase (Xc ) in the FA powders increases after the thermal annealing process. Because it is known that the crystallinity is inuenced by the crystallite size and lattice strain (Nakano et al., 2002), the investigation of structural features from broadened diffraction proles of samples enables us to calculate the crystallinity degree of samples. The presence of additional phases favors amorphization. These extra phases have also been found to affect the stability of the amorphous phases formed and structural features (crystallite size and lattice strain) of the product. According to Table 2, the amorphous characteristics are stronger for FA1 than for FA2, which corresponds to the presence of CaHPO4 as an extra phase. Similarly, the crystallinity degree is stronger for FA4 compared to FA3. In addition, the comparison between the crystallite size and lattice strain of samples before (FA1 and FA2) and after (FA3 and FA4) heat treatment conrms the thermal recovery of crystallinity at 600 C. Based on these results, we conclude that the impurity of CaHPO4 is an important parameter that affects the structural features of FA produced by the mechanochemical process.

where h, k, and l are the Miller indices of the reection planes. The (0 0 2) and (3 0 0) reections were chosen for the lattice parameter calculation (Qian, Kang, Zhang, & Li, 2008). Hexagonal lattice parameters and their variations for heattreated and unheated FA synthesized in this study are summarized in Table 3. Hexagonal lattice parameters a and c differed for all samples before and after heat treatment. These variations are more apparent for the unit cell parameter a, whereas the other parameter, c, was slightly altered with respect to a. Comparison between lattice parameters before and after heat treatment conrms the lattice distortion of FA during milling, while after the thermal annealing process at 600 C, recovery of crystallinity for FA3 and FA4 was observed. The increase of crystallinity degree as a result of the thermal annealing process is also conrmed by the values calculated using the Landi equation (Landi et al., 2000). As mentioned above, apatite phases with higher crystallinity have lower activity towards bioresorption and lower solubility in the physiological environment (Nakano et al., 2002; Seckler et al., 1999). A previous report (Mohammadi Zahrani & Fathi, 2009) reported that the size and number of balls had no signicant effect on the synthesis time and grain size of FA, while decreasing the rotation speed or ball to powder weight ratio increased the synthesis time and the grain size of FA. In addition, our experimental results conrm that the chemical composition of the initial materials and the thermal annealing process are important parameters that affect the crystallinity degree of products synthesized via the mechanochemical method. However, optimizing the mechanical alloying parameters to synthesize pure nanostructured FA that could satisfy the requirements of biomedical specications (mechanical properties and biocompatibility) must be carried out in future works.

Table 3 Lattice parameters of the samples in comparison with those of standard FA (JCPDS #15-0876) in A. Sample a-Axis c-Axis a c FA1 9.411 6.890 +0.043 +0.006 FA2 9.374 6.878 +0.006 0.006 FA3 9.358 6.890 0.010 +0.006 FA4 9.377 6.883 +0.009 0.001 Standard FA 9.368 6.884

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

541

3.2. FT-IR evaluation The vibrational spectroscopic methods provide information about the phosphate and hydroxide vibrations, which are very sensitive to substitution within the apatite lattice (Wei, Evans, Bostrom, & Grondahl, 2003); hence, FT-IR spectra were used to provide evidence for the identity of the apatite powders. The spectra shown in Fig. 3 reect the presence of functional groups that determine the composition of the samples and the changes that occurred during the mechanothermal process. In this section, the samples after annealing (FA3 and FA4) were evaluated using FT-IR spectroscopy because of the absence of polymeric contamination (as a result of vial wear) and the presence of suitable characteristics in these samples. In the spectrum of sample FA3 (Fig. 3(a)), two bands relating to the vibration of the adsorbed water in the apatite structure were detected (Fathi & Mohammadi Zahrani, 2009a, 2009b), where one was observed as a broad peak between 3700 and 2800 cm1 and the other as the widened band at 1645 cm1 (Fathi & Mohammadi Zahrani, 2009a). The characteristic peaks of the phosphate group, which had four distinct asymmetrical stretching vibration modes, namely, 1 (965 cm1 ), 2 (477 cm1 ), 3 (1099 and 1048 cm1 ), and 4 (605 and 574 cm1 ), appeared in the broad spectrum. The 1 1 and 2 vibration peaks were observed at 965 and at 477 cm , respectively. As a major peak of the phosphate group, the 3 vibration peak was observed in the region between 1099 and 1048 cm1 . The band between 605 and 574 cm1 showed the 4 vibration mode of the phosphate group. In this sample, the absence of OH

bands at 630 cm1 and the appearance of a band at 738 cm1 corresponding to the shifting OH liberation mode indicate the relatively complete transformation of HA into FA, and therefore, the synthesized product is almost pure FA (Barinov et al., 2004). According to the indications found in the literature (Barinov et al., 2004; Fathi & Mohammadi Zahrani, 2009a, 2009b; Mohammadi Zahrani & Fathi, 2009), the appearance of this band was caused by the increase of uorine ion content in the (OH , F ) chain of apatite with a predominant conguration of OH F. The FT-IR spectrum of sample FA4 is presented in Fig. 3(b). In this spectrum, similar bands relating to the vibration of the adsorbed water in the apatite structure are also revealed (Fathi & Mohammadi Zahrani, 2009a, 2009b). The asymmetrical stretching ( 3 ) and bending ( 4 ) modes of PO4 ion are detected at approximately 1098 (s) and 1040 (s), and 603 (b) and 570 (b) cm1 , respectively (Fathi & Mohammadi Zahrani, 2009a, 2009b; Mohammadi Zahrani & Fathi, 2009). In addition, a doublet at 1422 and 1456 cm1 corresponding to 3 and a band at 865 cm1 corresponding to 2 vibration mode of carbonated groups appeared. These peaks showed that FA contained some CO3 2 groups in PO4 3 sites of the apatite lattice (B-type substitution) (Lafon, Champion, & Bernache-Assollant, 2008). Because carbonates are constituents of hard tissue structures (Sanosh et al., 2009), the presence of low content of CO3 2 may improve the bioactivity of FA. In addition, for this sample, the band at 722 cm1 corresponds to the shifting OH liberation mode, which is caused by the increase of F content in the (OH , F ) chain of apatite (Barinov et al., 2004). It should be noted that the intermediate peaks at approximately 2929 and

Fig. 3. FT-IR spectra of mechanothermally treated samples of FA3 (a) and FA4 (b).

542

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

Fig. 4. TEM bright eld images of nanopowders after mechanochemical process (273 K, air atmosphere) of samples FA1 (a) and FA2 (b), showing the spheroidal morphology of nanoparticles.

2390 cm1 are not characteristic of the samples; these peaks belong to KBr itself (Nath, Tripathi, & Basu, 2009). Therefore, based on the FT-IR analysis, the synthesized products via the mechanochemical process have high uorine contents in apatite structure. It has been found (Nath et al., 2009) that nanocrystalline FA with high uorine content can be suitable for dental applications because these compounds are less soluble in acidic conditions. 3.3. Morphological and elemental analyses The size and morphology of ne powders may be determined with TEM. The bright eld transmission electron microscopic images of FA1 and FA2 samples are shown in Fig. 4. The FA1 particles are spheroidal with an average diameter of 25 5 nm (Fig. 4(a)). However, it should be noted that the particles do not possess high regularity in shape; in other words, their surfaces are not smooth.

In addition, the particles show a high tendency towards agglomeration. Fig. 4(b) shows that the sample FA2 possesses a mostly spherical structure with an average diameter of 31 6 nm. The SEM micrograph of the FA3 sample (Fig. 5(a)) indicate that the synthesized powder contained large agglomerates, which consist of signicantly ner particles and agglomerates with spheroidal morphology. This result is in agreement with other references (Barinov et al., 2004; Fathi & Mohammadi Zahrani, 2009a). In addition, the TEM observations conrm that the particles of sample FA3 are spheroidal with an average diameter of 30 7 nm (Fig. 5(b)). As shown in Fig. 5(c), the results of the measurement of the elemental composition conrm the presence of uorine in the FA3 sample. Fig. 6(a) presents SEM micrographs of the FA4 sample. According to Fig. 6(a), the chemical interactions at the contacting surface of particles resulted in substantial compactness of the particles and the formation of large agglomerates, which consist of ne particles

Fig. 5. Morphological and elemental analyses of the FA3 sample: (a) SEM image, (b) TEM image and (c) EDX spectrum.

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544

543

Fig. 6. Morphological and elemental analyses of the FA4 sample: (a) SEM image, (b) TEM image and (c) EDX spectrum.

with spheroidal morphology (see higher magnication). The TEM image of sample FA4 (Fig. 6(b)) shows the particles to be more spherical than the unheated FA2 particles, and the average size of the particles is 29 9 nm. As shown in the FT-IR spectra, some carbonate was detected in sample FA4; however, the absence of carbon and the presence of uorine in the small area EDX analysis conrm the high uorine/carbon ratio, which reveals the low amount of carbonated apatite compounds in sample FA4 (Fig. 6(c)). According to EDX analyses, there is a marked difference in the chemical compositions of FA3, FA4 and the empirical formula. These variations may be related to some impurities in the FA structure that affect the values of the Ca/P ratio (approximately 2.5). It was reported (Wei et al., 2003) that such high Ca/P ratios may be found in apatites containing additional calcium as calcium oxide or hydroxide impurities. However, the presence of uorine in small area EDX analysis conrms the presence of uorine in the FA3 and FA4 samples. The structural and morphological features of the samples in comparison with other references (Fathi & Mohammadi Zahrani, 2009a; Nasiri-Tabrizi, Honarmandi, Ebrahimi-Kahrizsangi, & Honarmandi, 2009) are presented in Table 4. According to Table 4, the maximum particle size measured with TEM is below the crystallite size calculated from the line broadening of the X-ray diffraction peak. Thus we determined that, after 60 h of milling

and subsequent thermal treatment at 600 C, this method gives rise to single-crystal FA with average sizes of 30, 37, 37 and 38 nm for FA1, FA2, FA3 and FA4, respectively. A novel mechanochemical process for synthesizing nanosize single-crystal FA with spheroidal shape was thus developed. In most reports concerning the synthesis of FA, the particle shapes are plates (Rameshbabu, Sampath Kumar, & Prasad Rao, 2006) or polyhedra (Barinov et al., 2004; Fathi & Mohammadi Zahrani, 2009a), but nanoparticles with spheroidal morphology were successfully prepared in the current research. Because the spherical geometry rather than irregular shape is important for achieving osseointegration (Hsu, Turner, & Miles, 2007; Nayar, Sinha, Basu, & Sinha, 2006), the products synthesized via mechanothermal processes are preferred for medical applications. It should be noted that the use of polymeric milling media has been suggested in the present study not only to annihilate contamination but also to achieve modied morphologies with good biomedical performance (Nasiri-Tabrizi et al., 2009). However, it was conrmed using SEM and TEM that the mechanochemically prepared powders have nanocrystalline structures with the appropriate morphology. The suitable structural and morphological features of mechanochemically synthesized powders and high efciency of this process presents a new way to produce commercial amounts of nanostructured FA powder. The comparison of the

Table 4 Comparison of structural and morphological features of the synthesized samples with other references. Sample FA1 FA2 FA3 FA4 FA HA Source This study This study This study This study Fathi and Mohammadi Zahrani (2009a) Nasiri-Tabrizi et al. (2009) DXRD (nm) 25 6 35 3 36 2 36 3 49 24 DTEM (nm) 25 5 31 6 30 7 29 9 35 17 8 Shape Spheroidal Spheroidal Spheroidal Spheroidal Polyhedron Rod Structure Low crystalline Crystalline High crystalline High crystalline High crystalline High crystalline

544

R. Ebrahimi-Kahrizsangi et al. / Particuology 9 (2011) 537544 Jantova, S., Theiszova, M., Letasiova, S., Birosova, L., & Palou, T. M. (2008). In vitro effects of uor-hydroxyapatite, uorapatite and hydroxyapatite on colony formation, DNA damage and mutagenicity. Mutation Research, 652, 139144. Kalita, S. J., Bhardwaj, A., & Bhatt, H. A. (2007). Nanocrystalline calcium phosphate ceramics in biomedical engineering. Materials Science and Engineering C, 27, 441449. Kurmaev, E. Z., Matsuya, S., Shin, S., Watanabe, M., Eguchi, R., Ishiwata, Y., et al. (2002). Observation of uorapatite formation under hydrolysis of tetracalcium phosphate in the presence of KF by means of soft X-ray emission and absorption spectroscopy. Journal of Materials Science: Materials in Medicine, 13, 3336. Lafon, J. P., Champion, E., & Bernache-Assollant, D. (2008). Processing of AB-type carbonated hydroxyapatite Ca10x (PO4 )6x (CO3 )x (OH)2x2y (CO3 )y ceramics with controlled composition. Journal of the European Ceramic Society, 28, 139147. Landi, E., Tampieri, A., Celotti, G., & Sprio, S. (2000). Densication behavior and mechanisms of synthetic hydroxyapatites. Journal of the European Ceramic Society, 20, 23772387. Mohammadi Zahrani, E., & Fathi, M. H. (2009). The effect of high-energy ball milling parameters on the preparation and characterization of uorapatite nanocrystalline powder. Ceramics International, 35, 23112323. Nakano, T., Tokumura, A., & Umakoshi, Y. (2002). Variation in crystallinity of hydroxyapatite and the related calcium phosphates by mechanical grinding and subsequent heat treatment. Metallurgical and Materials Transactions A, 33, 521528. Nasiri-Tabrizi, B., Honarmandi, P., Ebrahimi-Kahrizsangi, R., & Honarmandi, P. (2009). Synthesis of nanosize single-crystal hydroxyapatite via mechanochemical method. Materials Letters, 63, 543546. Nath, S., Tripathi, R., & Basu, B. (2009). Understanding phase stability, microstructure development and biocompatibility in calcium phosphatetitania composites, synthesized from hydroxyapatite and titanium powder mixture. Materials Science and Engineering C, 29, 97107. Nayar, S., Sinha, M. K., Basu, D., & Sinha, A. (2006). Synthesis and sintering of biomimetic hydroxyapatite nanoparticles for biomedical applications. Journal of Materials Science: Materials in Medicine, 17, 10631068. Nikcevic, I., Jokanovic, V., Mitric, M., Nedic, Z., Makovec, D., & Uskokovic, D. (2004). Mechanochemical synthesis of nanostructured uorapatite/uorhydroxyapatite and carbonated uorapatite/uorhydroxyapatite. Journal of Solid State Chemistry, 177, 25652574. Palmer, C. A., & Anderson, J. J. B. (2001). Position of the American dietetic association: The impact of uoride on health. Journal of the American Dietetic Association, 101, 126132. Qian, J., Kang, Y., Zhang, W., & Li, Z. (2008). Fabrication, chemical composition change and phase evolution of biomorphic hydroxyapatite. Journal of Materials Science: Materials in Medicine, 19, 33733383. Rameshbabu, N., Sampath Kumar, T. S., & Prasad Rao, K. (2006). Synthesis of nanocrystalline uorinated hydroxyapatite by microwave processing and its in vitro dissolution study. Bulletin of Material Science, 29, 611615. Rhee, S. H. (2002). Synthesis of hydroxyapatite via mechanochemical treatment. Biomaterials, 23, 11471152. Rodriguez-Lorenzo, L. M., Hart, J. N., & Gross, K. A. (2003). Inuence of uorine in the synthesis of apatites. Synthesis of solid solutions of hydroxy-uorapatite. Biomaterials, 24, 37773785. Roeder, R. K., Converse, G. L., Leng, H., & Yue, W. (2006). Kinetic effects on hydroxyapatite whiskers synthesized by the chelate decomposition method. Journal of American Ceramic Society, 89, 20962104. Sanosh, K. P., Chu, M. C., Balakrishnan, A., Lee, Y. J., Kim, T. N., & Cho, S. J. (2009). Synthesis of nano hydroxyapatite powder that simulate teeth particle morphology and composition. Current Applied Physics, 9, 14591462. Seckler, M. M., Danese, M., Derenzo, S., Valarelli, J. V., Giulietti, M., & RodriguezClemente, R. (1999). Inuence of process conditions on hydroxyapatite crystallinity obtained by direct crystallization. Materials Research, 2, 5962. Silva, C. C., Graca, M. P. F., Valente, M. A., & Sombra, A. S. B. (2007). Crystallite size study of nanocrystalline hydroxyapatite and ceramic system with titanium oxide obtained by dry ball milling. Journal of Materials Science, 42, 38513855. Silva, C. C., Pinheiro, A. G., Miranda, M. A. R., Goes, J. C., & Sombra, A. S. B. (2003). Structural properties of hydroxyapatite obtained by mechanosynthesis. Solid State Sciences, 5, 553558. Suryanarayana, C. (2001). Mechanical alloying and milling. Progress in Materials Science, 46, 1184. Wei, M., Evans, J. H., Bostrom, T., & Grondahl, L. (2003). Synthesis and characterization of hydroxyapatite, uoride-substituted hydroxyapatite and uorapatite. Journal of Materials Science: Materials in Medicine, 14, 311320. Zhang, H., Zhu, Q., & Xie, Z. (2005). Mechanochemicalhydrothermal synthesis and characterization of uoridated hydroxyapatite. Materials Research Bulletin, 40, 13261334.

mechanical properties and biocompatibility between other apatite structures and the novel nanostructures will be carried out in future work. 4. Conclusions In summary, FA nanostructures were successfully prepared by novel and simple mechanochemical processes. The XRD analysis reveals that the chemical composition of the initial materials and the thermal annealing process are important parameters in the mechanochemical synthesis of FA with high crystallinity. The FT-IR studies performed after annealing show that all the characteristic bands corresponding to FA appear in samples FA3 and FA4. Further, the synthesized products have various amounts of uorine in the apatite structure, which has potential for biomedical applications. According to the EDX spectra, the absence of carbon in a small area of analysis conrms the low amount of carbonated apatite compounds in sample FA4. Morphological evaluations showed that the annealing at 600 C does not accompany a change in the morphology of the powders and that the outcome morphology is spheroidal. In addition, the single-crystal FA powder exhibits an average size of 3038 nm. The results conrm that the mechanothermal process suggests a new method to produce commercial amounts of nanostructured FA powder with high quality and suitable structural and morphological features. Acknowledgements The authors are grateful to the Research Affairs of Islamic Azad University and to Najafabad Branch for supporting this research. References
Barinov, S. M., Shvorneva, L. I., Ferro, D., Fadeeva, I. V., & Tumanov, S. V. (2004). Solid solution formation at the sintering of hydroxyapatiteuorapatite ceramics. Science and Technology of Advanced Materials, 5, 537541. Chen, Y., & Miao, X. (2005). Thermal and chemical stability of uorohydroxyapatite ceramics with different uorine contents. Biomaterials, 26, 12051210. Cheng, K., Zhang, S., & Weng, W. (2006). Solgel preparation of uoridated hydroxyapatite in Ca(NO3 )2 PO(OH)3x (OEt)x HPF6 system. Journal of Sol-Gel Science and Technology, 38, 1317. De Castro, C. L., & Mitchell, B. S. (2002). Synthesis functionalization and surface treatment of nanoparticles. In M. I. Baraton (Ed.), Nanoparticles from mechanical attrition (pp. 114). Stevenson Ranch, CA: American Scientic Publishers. El Briak-BenAbdeslam, H., Ginebra, M. P., Vert, M., & Boudeville, P. (2008). Wet or dry mechanochemical synthesis of calcium phosphates? Inuence of the water content on DCPDCaO reaction kinetics. Acta Biomaterialia, 4, 378386. Fathi, M. H., & Hani, A. (2007). Evaluation and characterization of nanostructure hydroxyapatite powder prepared by simple solgel method. Materials Letters, 61, 39783983. Fathi, M. H., & Mohammadi Zahrani, E. (2009a). Fabrication and characterization of uoridated hydroxyapatite nanopowders via mechanical alloying. Journal of Alloys and Compounds, 475, 408414. Fathi, M. H., & Mohammadi Zahrani, E. (2009b). Mechanical alloying synthesis and bioactivity evaluation of nanocrystalline uoridated hydroxyapatite. Journal of Crystal Growth, 311, 13921403. Fini, M., Savarino, L., Nicoli Aldini, N., Martini, L., Giavaresi, G., Rizzi, G., et al. (2003). Biomechanical and histomorphometric investigations on two morphologically differing titanium surfaces with and without uorohydroxyapatite coating: An experimental study in sheep tibiae. Biomaterials, 24, 31833192. Hsu, Y. H., Turner, I. G., & Miles, A. W. (2007). Fabrication and mechanical testing of porous calcium phosphate bioceramic granules. Journal of Materials Science: Materials in Medicine, 18, 19311937.

You might also like