You are on page 1of 5

Materials Letters 60 (2006) 3227 – 3231

www.elsevier.com/locate/matlet

Surfactant-assisted synthesis of hydroxyapatite particles


YingJun Wang a,b , JingDi Chen a,b,⁎, Kun Wei a,b , ShuHua Zhang a,b , XiDong Wang a,b
a
Institute of Materials, South China University of Technology, PR China
b
Key Laboratory of Specially Functional Materials and Advanced Manufacturing Technology of Ministry of Education, Guangzhou 510640, PR China

Received 27 October 2005; accepted 23 February 2006


Available online 20 March 2006

Abstract

Various morphologies of nano hydroxyapatite particles (HAP; Ca10(PO4)(OH)2) were synthesized through the hydrothermal method using
CTAB as a template. TEM micrographs reveal the hydroxyapatite particles ranged from spheroid to long fibers. The obtained smallest average
diameter of spheroidal particles, ∼27 nm, were achieved at 90 °C, pH = 13 for 20 h. The biggest long fibers, more than 1125 nm and ca. 60 nm in
width, were obtained at 150 °C, pH = 9 for 20 h. The morphology of the obtained particles presented a certain rule with the temperature at different

pH values. A probable mechanism involves the interaction of the tetrahedral CTAB cation structure, the tetrahedral PO3− 4 structure and the OH
ions at different pH values. The results show that the nano hydroxyapatite particles can be tailored through the control of the process temperature,
the reaction time, the concentration of CTAB and the pH value of the solution.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Hydroxyapatite; Nanoparticle; Hydrothermal

1. Introduction nanosize HAP can be exploited in many other new fields.


Although ultrafine, non-aggregated calcium phosphate nanopar-
Hydroxyapatite (HAP) is the principal inorganic constituent of ticles have been previously synthesized by microemulsion based
bones and teeth [1]. Synthetic HAP crystals are usually prepared routes, which were in the form of non-aqueous dispersions [10].
as rods or needles, which are similar in structure and composition However, the common aqueous synthetic route, the hydrothermal
to HAP found in human bone. The importance of HAP has led to method, which has proved to be a convenient way to prepare
extensive research in numerous areas ranging from the physico- materials, including salts and metal oxides, has also been applied
chemical mechanisms of the formation to its applicability as a to prepare nanosized HAP particles, but with poor control on
biomedical or industrial material [2–5]. In particular, the morphology [11]. Although the control over microstructure seems
biocompatibility and osteoconductive properties of HAP have too big a challenge with traditional methods, the biological
made it useful as implant material. New developments on the process itself has provided some clues to achieve this: the
production of nanosized HAP particles have led to many new controlled nucleation and crystal growth process mediated by
applications. For example, nanosized HAP particles can retard the macromolecular control and cell organization would finally result
multiplication of cancer cells and be used as an efficient drug in uniform products. In the presence of extraneous additives, the
delivery agent [6,7]. role of crystal surface is more analogous to the conventional view
The specific characteristics, crystallization process and of host–guest systems. The surface layers of the crystal can
bioactivity of diverse nano HAP crystals have led to different incorporate soluble additives provided that there is a degree of
clinical application [8,9]. If the crystal can be presented not only complementarity in charge and size between the guest ions and
as rods and needles but also in spherical and flaky shapes, the the interstices in the structure of the crystal boundary layers.
Several macromolecules, such as stearic acid, monosaccharides
⁎ Corresponding author. Institute of Materials, South China University of and related molecules have been explored with desired control on
Technology, PR China. the morphology. Cetyltrimethylammonium bromide (CTAB), as
E-mail address: ccjjdd@21cn.com (J. Chen). one of the macromolecules, is also widely used to achieve the
0167-577X/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.matlet.2006.02.077
3228 Y. Wang et al. / Materials Letters 60 (2006) 3227–3231

control of the morphology in many aqueous synthetic methods fraction Data), where intermediates (e.g. tricalcium phosphate) are only
[12–14]. present in one sample obtained at 150 °C for 20 h when pH was 9.
Here we report the synthesis of HAP nanoparticles through the Fig. 5a and b show the FTIR spectra of the samples synthesized at
hydrothermal method with controlled size and shape by using the 150 °C, pH = 13 for 20 h (Fig. 5a) and 120 °C, pH = 9 for 20 h (Fig. 5b)
respectively (other data not shown). The 470 cm− 1 band results from
water soluble macromolecule surfactant, cetyltrimethylammo-
nium bromide (CTAB). The influence of temperature, pH of the
solution and surfactant on the precipitated phases and their
crystallinity of the products are analysed.

2. Experimental

HAP was synthesized using CTAB as surfactant and calcium


chloride (CaCl2) and phosphoric acid (H3PO4, 85%) as calcium
and phosphorous sources, respectively. 1.729 g of phosphoric
acid and 4.370 g CTAB were dissolved in 50 ml de-ionized
water. The pH was adjusted to a value between 9 and 13 using
sodium hydroxide. 2.775 g of CaCl2 was dissolved into 30 ml of
de-ionized water. Subsequently, the CaCl2 solution was added
dropwise into the phosphorous-surfactant solution mixed under
vigorous stirring, yielding a milky suspension. The suspension
was stirred for half an hour. It was then transferred into a 100 ml
PTFE-lined autoclave and heated at 90–150 °C in an oven for a
specified time. The precipitate was then filtered off and washed
with de-ionized water. A gel-like paste was obtained, which was
dried in an oven at 100 °C for 24 h. The powder was then
calcined in a furnace at 550 °C for 6 h, yielding a white powder.
The sample structure was analysed by X-ray diffraction
(XRD), carried out with a PANalytical X′Pert PRO X-ray
diffractometer with Cu Kα (λ = 0.15418 nm) incident radiation.
The morphology of the products was studied by Transmission
Electron Microscopy (TEM, TECHAI G2 12 and Philips
CM300). Fourier transform IR spectra were collected on a
Nicolet Nexus spectrometer by using a KBr wafer technique in
order to study the composition of the samples.

3. Results and discussion

From the TEM micrograph shown in Fig. 1, we can observe most


particles present nearly round shape with the average diameter of
27 nm at 90 °C for 20 h when pH = 13 in Fig. 1a. When the temperature
goes up to 120 and 150 °C, the morphology of the obtained HAP
crystals is similar to the Fig. 1a with the average diameter of 74 nm
(Fig. 1b) and 85 nm (Fig. 1c). There are also few crystals with rod-like
in Fig. 1. The TEM image of Fig. 2a shows that a majority of needles
with average diameter of 27 nm and 255 nm in length precipitated at
90 °C for 20 h when pH = 9. With the rising temperature, we obtained
the mainly elongated fibers with average length up to 820 and 34 nm in
diameter (Fig. 2b) and long fibers with average diameter of 60 and
1125 nm in length (Fig. 2c). When pH decreases from 13 to 9, the
morphology of HAP changed from nearly round shape (Fig. 1) to
needles (Fig. 2a) and long fibers (Fig. 2b and c). We can also see that
the average diameters (both in Figs. 1 and 2) and the length (in Fig. 2)
of the synthesized crystals are increasing with the rising temperature.
Figs. 3 and 4 are the X-ray diffraction patterns corresponding to
Figs. 1 and 2, respectively. The diffraction patterns revealed
characteristic peaks of HAP after calcination, where the filled squares
correspond to the standard characteristic peaks for HAP. We observed
high consistency between the data from our samples and that obtained Fig. 1. TEM micrograph of HAP precipitated at different temperatures for 20 h
from the standard data (1997, JCPDS-International Center for Dif- when pH = 13: (a) 90 °C, (b) 120 °C, (c) 150 °C.
Y. Wang et al. / Materials Letters 60 (2006) 3227–3231 3229

the γ2 phosphate mode [15]. The bands at 560, 600 cm− 1 are derived
from the γ4 phosphate modes. The peak at 960 cm− 1 is assigned to γ1
phosphate mode. The bands at 1039 and 1086 cm− 1 stand for the γ3
phosphate mode [16]. The absorption band at 1639 cm− 1 reflects H2O
bending mode. The band at 3439 cm− 1 band may come from lattice
H2O because this band exists in the range of 3550–3200 cm− 1. The
stretching vibration and bending modes of the OH− appeared at 3569
and 631 cm− 1, respectively [17]. The weak absorption bands at 1408,

Fig. 3. The XRD patterns of the HAP precipitated at 90 °C (a), 120 °C (b) and
150 °C (c) for 20 h when pH = 13.

1456 cm− 1 are observed, those reported for carbonated HAP [18,19].
The existence of carbonate phosphate might come from the atmosphere
carbon dioxide while dissolving, stirring, reaction and the calcining
processes. The highly sensitive FTIR results indicate that no CTAB
molecule is incorporated in the HAPs [14].
Through the XRD and FT-IR, it proved that the synthesized crystals
were all HAP. The samples were kept in the same autoclave for the
same time, but the morphologies were quite different. Some of them
were also different from the most reported samples prepared by the
hydrothermal methods whose crystalline morphology is usually rod or
needle-like. This implies that the different experimental condition such
as surfactant, pH and temperature might exert important effects on the
synthesis of HAP under the same reaction time.
According to the kinetics of crystallization [20], the lower
temperature results in the smaller critical sizes of nuclei and it is
easy for particles to reach the critical sizes. The lower the temperature
is, the slower the growth rate of crystal is. These induce the multi-small
ultimate products at low temperature. But with the rising temperature,
the critical size of nuclei is also increasing. Owing to the enormous
surface energy, the small particles dissolve again and cannot grow to
stable nuclei. This makes few crystals beyond critical size of nuclei.

Fig. 2. TEM micrograph of HAP precipitated at different temperatures for 20 h Fig. 4. The XRD patterns of the HAP precipitated at 90 °C (a), 120 °C (b) and
when pH = 9: (a) 90 °C, (b) 120 °C, (c) 150 °C. 150 °C (c) for 20 h when pH = 9.
3230 Y. Wang et al. / Materials Letters 60 (2006) 3227–3231

Fig. 7. Influence of the temperature on the average length of HAP crystals


Fig. 5. Typical FT-IR spectra of HAP: (a) at 150 °C, pH = 13 for 20 h; (b) at precipitated at pH = 9 for 20 h.
120 °C, pH = 9 for 20 h.

correlate with the charge and stereochemistry properties. In an aqueous


Besides, the crystals grow fast at high temperature. The higher tem- system, CTAB would ionize completely and result in a cation with
perature will result in bigger products. Because the hydrothermal tetrahedral structure. Meanwhile, the phosphate anion is also a
method is a close system, increasing temperature will make the tetrahedral structure. It is then proposed that the charge and structure
pressure rise and the supersaturation decrease. This is also very complementarity endows CTAB with the capability to control the
advantageous for crystals to grow with the rising temperature. So the crystallization process [13]. But when pH is higher, there are many
crystals will grow bigger with rising temperature. Fig. 6 shows the OH− ions in the aqueous solution. Although the charge of OH− is lower

average diameter of the samples synthesized at different temperature than that of PO3−
4 , the concentration of OH is higher than that of PO4
3−

for 20 h when pH = 13 and pH = 9 (drawn from Figs. 1 and 2). The and the diameter of OH− is smaller than that of PO3− 4 . Moreover, the
length of HAP crystals precipitated at different temperature for 20 h repulsive interaction of OH− and PO3− 4 decreases the colliding
when pH = 9 is given in Fig. 7 (drawn from Fig. 2). From Figs. 6 and 7, probability of PO3− 4 and CTAB tetrahedral structures and prevent the
we find that both of the diameter and length are increasing with the interaction of the two tetrahedral structures. So it disturbs the HAP
rising temperature. But when we compare the Figs. 1 and 2, we can see crystal to crystallize as nanorods. While the pH is lower, the OH− ions
that the morphology of the synthesized HAP crystals is quite different are fewer. The morphology of the HAP crystal would be still rod-like.
even at the same temperature. So it suggests that the adding surfactant, Fig. 8 shows the effects between the surfactant cation, phosphate anion
CTAB, and pH might be the main reason of this phenomenon. and hydroxyl.
The HA monocrystalline has the trend to grow following the c-axis Crystallization is a complicated process, which is affected by
of the HA structure [21]. The previous study of hydrothermal synthesis thermodynamics, kinetics and has relationship with dissolution,
of HAP showed that the morphology of HAP would be nanorods using temperature, supersaturation, stirring and impurity etc. [20]. At the
CTAB as surfactant [13]. The behavior of CTAB was considered to early age of synthesizing HAP in the aqueous system, it may result in
ACP (amorphous calcium phosphate) type and OCP (octacalcium
phosphate) type precursors, then transformate into HAP [22]. The
transformation speed is depended on pH. The conversion rate increases
with the rising pH while pH values are between 7 and 10. But it
decreases with the rising pH while pH beyond 10 [23]. Meanwhile, the
OH− ions do not destroy the interaction of the tetrahedral CTAB cation
structure and the tetrahedral PO3−4 structure when pH is 9 because of the
low OH− concentration. So it promotes the HAP crystal growing along
c axis and resulted in the needles and elongated fibers in Fig. 2. When

Fig. 6. Influence of the temperature on the average diameter of HAP crystals Fig. 8. A schematic drawing showing the effects between the surfactant cation,
obtained at pH = 13 (a), pH = 9 (b) for 20 h. phosphate anion and hydroxyl.
Y. Wang et al. / Materials Letters 60 (2006) 3227–3231 3231

pH is 13, the OH− concentration is so high that it destroys the NO. [2005]4). Prof. Dr. Betty Leon (Applied Physics Depart-
interaction of the tetrahedral CTAB cation and the tetrahedral PO3− 4 ment, University of Vigo, Spain) gave some good suggestions to
structures. Besides, the OH− ions also repulses the PO3− 4 ions. These this paper.
influences exert more negative effects on HAP crystallization. So it
prevents the HAP crystals growing along c axis. This is proved by the
nearly round grains in Fig. 1. However, with the proceeding of the References
reaction in the aqueous system, the gradual synthesis of HAP made the
OH− ions decrease. This also decreased the negative effect. But at that [1] S.V. Dorozhkin, M. Epple, Angew. Chem., Int. Ed. Engl. 41 (2002)
− 3130–3146.
time, the concentrations of PO3− 4 and OH were so low that few
[2] S. Koutsopoulos, E. Dalas, Langmuir 16 (2000) 6739.
nanorods of HAP still appeared in Fig. 1. [3] S. Koutsopoulos, E. Dalas, J. Colloid Interface Sci. 231 (2000) 207.
[4] S. Koutsopoulos, E. Dalas, Langmuir 17 (2001) 1074.
[5] H. Tanaka, K. Miyajima, M. Nakagaki, S. Shimabayashi, Chem. Pharm.
4. Conclusions Bull. 37 (1989) 2897.
[6] Y.F. Wang, Y.H. Yan, W. Ren, Trans. Nonferr. Met. Soc. China 14 (2004)
With adding adequate CTAB, it has been shown that the size 29–32.
[7] T.Y. Lu, S.Y. Chen, D.W. Liu, S.C. Liou, J. Control. Release 101 (2005)
and morphology of precipitated hydroxyapatites can be affected 112–121.
by control of the precipitation temperature and pH. Our results [8] K. de Groot, Biomaterials 1 (1980) 47.
show that this hydrothermal technique is particularly well suited [9] L.L. Hench, J. Am. Ceram. Soc. 74 (1991) 1487.
to control HAP size and morphology through variation of both [10] D. Walsh, J.L. Kingston, B.R. Heywood, S. Mann, J. Cryst. Growth 133
thermodynamic and nonthermodynamic processing variables (1993) 1.
[11] M. Yoshimura, H. Suda, K. Okamoto, K. Ioku, J. Mater. Sci. 29 (1994)
mentioned above. A probable mechanism for the controllable 399–402.
size and morphology is the interaction of CTAB, PO43− and OH− [12] M.H. Cao, Y.H. Wang, Langmuir 20 (2004) 4784–4786.
mixtures. At low OH− concentration, CTAB–PO43− exert most [13] Y. Li, Y.D. Li, Z.X. Deng, Int. J. Inorg. Mater. 3 (2001) 633–637.
effects on morphology. But at higher OH− concentration, OH− [14] Y.K. Liu, Y.Z. Wang, Mater. Lett. 56 (2002) 496–501.
[15] B.O. Fowler, Structural Properties of Hydroxyapatites and Related
ions interact with CTAB and PO43− and take negative effects on
Compound, Gaithersburg, 1968.
morphology. These result in the different morphology of HAPs. [16] I.R. Gibson, S.M. Best, J. Biomed. Mater. Res. 44 (1999) 422–428.
[17] T. Ishikawa, M. Wakamura, Langmuir 5 (1989) 140.
Acknowledgements [18] K. Ioku, M. Yoshimura, Nippon Kagaku Kaishi (1985) 1565.
[19] D.W. Halcomb, R.A. Young, Tissue Int. 31 (1980) 189.
We acknowledge the financial support from National Natural [20] J.W. Mullin, Crystallization, Butterworth-Heinemann, 1997.
[21] E. Bouyer, F. Gitzhofre, J. Mater Sci., Mater. Med. 11 (2000) 523–531.
Science Foundation of China (Grant No. 50572029, 50272021 [22] R. Rodriguez-Clemente, A. Lbpez-Macipe, J. Eur. Ceram. Soc. 18 (1998)
and 50472054), Natural Science Foundation Team Project of 1351–1356.
Guangdong, China (Grant No. 04205786), Key Science and [23] J.L. Meyer, C.C. Weatherall, J. Colloid Interface Sci. 89 (1982) 257–267.
Technology Project of the Chinese Ministry of Education (Grant

You might also like