Materials Sciences and Applicatio ns, 2011, 2, 105-110
doi:10.4236/msa.2011.22014 Published Online February 2011 (http://www.SciRP.org/journal/msa)
Copyright © 2011 SciRes. MSA
105
Effect of Hydroxide Ion Concentration on the
Morphology of the Hydroxyapatite Nanorods
Synthesized Using Electrophoretic Deposition
Khalil Abdelrazek Khalil, Abdulhakim A. Almajid, Mahmoud S. Soliman
Mechanical Engineering Department, College of Engineering, King Saud University, Riyadh, Saudi Arabia.
Email: kabdelmawgoud@ksu.edu.sa, khalil305@hotmail.com
Received December 16th, 2010; revised January 5th, 2011; accepted January 23rd, 2011.
ABSTRACT
The effect of OH concentration on the morphology of the hydroxyapatite (HA) nanorod synthesized using electropho-
retic deposition (EPD) method has been investigated. The growth of HA nanorods was achieved on polished titanium
substrates. The electrolyte used in this study was prepared by dissolving calcium acetate (Ca (CH3COO)2 H2O), and
Ammonium dihydrogen phosphate (NH4H2PO4) in distilled water without any surfactant, and was maintained at 80-
130˚C. Two electrolytes with OH concentration of 104 and 1010 were prepared. A highly homogeneous HA nanorods
deposited on the titanium substrates were obtained after 1 h in the electrolytes with higher OH concentration of 104.
On the other hand, a flower-shaped HA nanostructures composed of needle-like HA crystals were obtained in the elec-
trolyte of lower OH concentration of 1010. The deposits were identified as HA crystal rods grown along the c axis and
perpendicular to the substrate. The HA deposits were characterized by scanning electron microscopy (SEM) while de-
tailed structural characterization was done using a transmission electron microscope (TEM) equipped with selected
area electron diffraction (SAED) patterns.
Keywords: Hydroxyapatite, Nanorods, Flower-Shaped, Electrophoretic Deposition, Synthesis
1. Introduction
Hydroxyapatite (HA: Ca10(PO4)6(OH)2) is a particularly
attractive material for human tissue implantation. HA has
similar chemical composition and crystal structure to
apatite in the human skeletal system and is therefore
suitable for bone substitution and reconstruction [1]. Hy-
droxyapatite can be synthesized via numerous production
routes, using a range of different reactants, such as solid-
state reactions [2], and wet chemical routes [3] based on
precipitation at low temperatures. These conventional
methods, however, usually prepare irregular forms of
powders. A hydrothermal method, which has proved to
be a convenient way to prepare materials including salts
and metal oxides, has also been applied, but control of
the morphology was poor [4]. Nevertheless, the size and
morphology largely determine the behavior of a certain
material, which is why a biomineralization process usu-
ally involves complicated mediation and the final prod-
ucts generally have a delicate microstructure [5]. Re-
cently there have been some reports of attempts to as-
semble the one-dimensional nanowires and nanorods di-
rectly into organized superstructures with the assistance
of surfactants. Kim et al. [6] explored the organization of
barium chromate nanorods at the water-air interface us-
ing the Langmuir-Blodgett (LB) technique. The inor-
ganic nanorods were successfully assembled into iso-
tropic, nematic, and smectic phases depending on the
surface pressure. Kasuga and coworkers have developed
a route to synthesize HA needle-like particles or fibers
using crystalline
-Ca3(PO4)2 fibers [7], but control of
the morphology was poor. Therefore, synthesis of high
crystalline HA nanorods is an urgent need of great sig-
nificance, because bone itself is a composite consisting
of HA nanorods embedded in the collagen matrix [8].
The application of surfactants as reverse micelles or mi-
croemulsions for the synthesis and self-assembly of
nanoscale structures is one of the most widely adopted
methods reported in the literature [9-18]. However, to the
best of our knowledge no studies have been reported on
the mechanism of fabrication of highly uniform HA
nanorods without a surfactant. In this paper, we report
Effect of Hydroxide Ion Concentration on the Morphology of the Hydroxyapatite Nanorods Synthesized Using
106
Electrophoretic Deposition
the effect of OH concentration on the morphology of the
hydroxyapatite (HA) nanorod synthesized using electro-
phoretic deposition (EPD) method. A simple and con-
trollable synthesis, without any surfactants, for highly
uniform HA nanorods based on the OH concentration
have been introduced. Two electrolytes with different
OH concentration, based on different pH values have
been used. The morphologies and structures of HA
nanorods were characterized with scanning electron mi-
croscopy (SEM) and transmission electron microscopy
(TEM).
2. Materials and Methods
Hydroxyapatite films were fabricated on polished tita-
nium substrates using the electrophoretic deposition (EPD)
method. A characteristic feature of this process is that
colloidal particles suspended in a liquid medium migrate
under the influence of an electric field (electrophoresis)
and are deposited onto an electrode. A titanium plate (10 ×
20 × 1 mm) was used as the cathode while a platinum
plate (20 × 20 × 0.5 mm) was used as the anode. The
electrolyte used in this study was prepared by dissolving
calcium acetate (Ca(CH3COO)2·H2O), and Ammonium
dihydrogen phosphate (NH4H2PO4) in distilled water.
The Ca/P ratio was adjusted to be 1.67, mimicking that
of natural hydroxyapatite. The solution was buffered to a
different pH value by ammonia. Calcium acetate and
Ammonium dihydrogen phosphate solutions were ini-
tially prepared separately, then mixing of the two elec-
trolytic solutions and the heating process were started
simultaneously to avoid excessive bulk precipitation.
Two electrolytes with OH concentration of 104 and 1010
were used, corresponding to the hydrogen ion [H+] con-
centration of 1010 and 104, based on the following equa-
tion:
14
w
HOH10K

 

  (1)
The corresponding pH values of 10 and 4, respectively,
were calculated as follows: Since the pH value is defined
as:
logpHH

(2)
The values of pH for the two solutions were 10 and 4,
respectively. The electrolyte was prepared in a Teflon
capped air-tight glass bottle with the electrode assembly.
The assembly (with electrolyte and electrodes) was
heated to a temperature ranging from 80 to 130˚C using a
silicon oil bath. During heating and synthesis, the solu-
tion was agitated with magnetic stirring at a constant
speed. Continuous stirring and aging are usually carried
out after the reactants have been combined while the cal-
cium is slowly incorporated into the apatitic structure.
This process also helps the material to approach stoi-
chiometric Ca/P ratio. The current was maintained at 12.5
mA/cm2 for 1 hour using a DC power supply. After elec-
trodeposition, the titanium plates were rinsed with dis-
tilled water and dried in air. General morphology of the
deposited HA nanorods was observed using scanning
electron microscopy (SEM, JEOL model JSM-6610LV,
Japan) and field-emission scanning electron microscope
(FE-SEM, JEOL model, Japan), while detailed structural
characterization was done with a transmission electron
microscope (TEM, JEOL JEM-2010, Japan) equipped
with selected area electron diffraction (SAED) patterns.
The SEM images were carried out by using a scanning
electron microscope with an energy dispersive X-ray
analyzer attached. Information about the phase and crys-
tallinity was obtained by using Rigaku X-ray diffracto-
meter (XRD, Rigaku, Japan) with Cu Kα (λ = 1.540 Å)
radiation over Bragg angle ranging from 10 to 80˚. Spec-
troscopic analysis of the grown HAp microcrystals was
carried out by FT-IR (AVATAR) using KBr pellet tech-
nique.
3. Results and Discussion
Hydroxyapatite nanorods deposited on the titanium sub-
strates were obtained in the electrolytes with OH con-
centration of 104 of a pH value of 10, after 1 hour of
electrophoretic processing at different electrolyte tempe-
ratures in the range of 80-130˚C, as shown in Figure 1.
The SEM observations show that the HAp crystals de-
posited from the higher concentration electrolyte (104
OH) are rod-like with a hexagonal cross section and
diameters of about 50-200 nm. The size of the deposited
nanorods increased remarkably with the electrolyte tem-
perature as shown in Figure 1.
Figure 2(a) shows FE-SEM image of HA nanorods. A
schematic representation of a single HA nanorod show-
ing the growth direction is shown in Figure 2(b). Se-
lected area electron diffraction (SAED) pattern image
obtained from the single HA nanorod projected along the
C-axis is shown in Figure 2(c). The schematic represent-
tation of a single HA nanorod suggested that, the HAp
crystal with the longitude (vertical) direction of [002], and
the six surface of the rod are (100). The size and mor-
phology of the as-synthesized hydroxyapatite products
were further examined by TEM and by selected area
electron diffraction (SAED) patterns (shown in Figure
2(c)). The diameters of most nanorods are almost the
same throughout their length, and all exhibited smooth
and clean surfaces.
There are numerous ways to grow crystals. The choice
of method depends greatly upon the physical and chemi-
Copyright © 2011 SciRes. MSA
Effect of Hydroxide Ion Concentration on the Morphology of the Hydroxyapatite Nanorods Synthesized Using 107
Electrophoretic Deposition
(a) (b)
(c) (d)
(e) (f)
Figure 1. FE-SEM micrographs of the hydroxyapatite nano-
rods deposited on the titanium substrates at (a) 80˚, (b) 90˚,
(c) 100˚, (d) 110˚, (e) 120˚, and (f) 130˚C.
cal properties of the sample. For solution methods of
crystallization, the solubility of the sample in various
solvent systems must be explored. If heating methods are
selected for growing crystals, the thermal stability and
melting point of the sample should be determined. Few
experiments were done to understand the growth process
of hydroxyapatite nanorods in the presence of much
higher OH concentration and based on that experiments
and results; we suggest here the formation mechanism
for the obtained products are as follows: the hydroxyapa-
tite is a polar hexagonal and highly anisotropic crystal
and generally grows along the c-axis direction. Due to
the anisotropic structure of hydroxyapatite (HAp), the
nanoparticles of HAp start to combine into one unit and
would be self-organized with each other and grow in
their most favorable crystallographic direction i.e. c-axis
[19]. In this accumulation process of the hydroxyapatite
nanoparticles, the presence of TiO2 (in the surface of the
titanium plate) is significant and it plays an important
role for the formation of well faceted hydroxyapatite
nanorods. Lee et al. [20] reported that the TiO2 is an ef-
fective nucleating agent for the calcium phosphates.
Figure 2. (a) FE-SEM image of HA nanorods and (b) A
schematic representation of a single HA nanorod showing
the growth direction, (c) Selected area electron diffraction
(SAED) pattern image obtained from the single HA nano-
rod projected along the C-axis.
Hence, one can propose that the TiO2 in this synthesis
acting as a catalyst to obtain a well crystallized c-axis
oriented hydroxyapatite nanorods.
On the other hand, the possible nanorod nucleation and
growth could be attributed to the relative specific surface
energies associated with the different planes of HA crystal
or nucleus. It is known that the crystal shape is deter-
mined by the relative specific surface energies associated
with the facets of this particular crystal. Hence, some
facets of the crystal have a preference to absorb OH ions
due to the different surface energies of the crystallite
facets, prior to the growth units are incorporated into a
crystal lattice, and this possibly restricts the movement of
Ca2+ and PO43 in the formed nucleus in one particular
direction. In this condition, OH templates the nucleation
process where free Ca2+ and PO43 react with OH in
uniaxial direction (planes with less OH concentration),
nucleating into HA nanorods. Thus, the shielding effect
of the OH ions on the interface will control the growth
rate of the OH absorbed facets [21]. Different facets of
the crystals have different quantity of the OH ions and
its hindrance effect. If the larger quantity of the OH will
be available at the interface, the hindrance effect of the
OH ions will be stronger on the facets which will affect
Copyright © 2011 SciRes. MSA
Effect of Hydroxide Ion Concentration on the Morphology of the Hydroxyapatite Nanorods Synthesized Using
108
Electrophoretic Deposition
the growth rate in the various crystal facets. As the syn-
thesis of the HAp nanorods was done at pH 10 (higher
OH concentration), HAp nanoparticles may generate
active site due to the strong absorption of the OH ions.
Moreover, the HAp crystals have faster growth rate along
the c-axis, i.e. (001) direction as compared to other
growth facets [22]. To know the role of higher OH con-
centration and TiO2, various experiments are underway.
Clearly more studies are required to obtain more con-
clusive evidences regarding the detailed growth process
for the formation HAp nanorods in the presence of higher
OH concentration and TiO2.
These HAp nanoparticles with OH ions in the pres-
ence of TiO2 may act as nuclei for the formation of HAp
nanorods. So the subsequent attachment and merging of
adjacent nanoparticles in accordance with the crystal
lattice of the HAp give rise to the phase transformation
from HAp nanoparticles to the HAp nanorods in the
presence of higher OH concentration and TiO2. Even
one can predict that the TiO2 is acting as a catalyst for the
transformation of HAp nanoparticles to the HAp nano-
rods but the actual role of higher OH concentration and
TiO2 are unclear. Remarkably, the size of the deposits
increased with increasing the electrolyte temperature.
The growth of HA nanorods with temperature can be
described by “oriented attachment” mechanism reported
for other ceramics like TiO2 [23].
In the present work, the electrolyte with much higher
OH concentration was employed and the rod-like HAp
crystal has a much smaller diameter. The smaller size of
the HAp crystal may be beneficial to the bioactivity of the
coating. All the deposits were rod-like in shape, having a
defined hexagonal crystal habit. Furthermore, it was found
that the long rod-like crystals grew up perpendicular to
the substrate. It is interesting to note that the synthesized
HAp nanorods reported in this paper are also grown in
c-axis at (001) direction as evident from the XRD data
(Figure 3). Hydroxyapatite nanorods deposited on the
titanium substrates were obtained in the electrolytes with
OH concentration of 1010, based on the pH value of 10,
after 1 h of electrophoretic processing at different elec-
trolyte temperatures as shown in Figure 1. The SEM
observation shows that the HAp crystals prepared with
higher concentration electrolyte (1 × 104 OH) are rod-
like with a hexagonal cross section and diameters of about
50-200 nm.
In the electrolyte of lower OH concentration of 1010,
a flower-shaped hydroxyapatite nanostructures composed
of needle-like HA crystal was obtained. Figure 3 shows
the SEM images of flower-shaped hydroxyapatite nanos-
tructures composed of nanorods grown on Ti substrate by
electrophoretic deposition at (a) 80˚C and (b) 100˚C (Fi-
(a) (a1)
(b) (b1)
Figure 3. SEM images of flower-shaped hydroxyapatite na-
nostructures composed of nanorods grown on Ti substrate
by electrophoretic deposition at (a) 80˚C and (b) 100˚C (a1
and b1 are high magnification images).
gures 3(a1) and 3(b1) are were taken at high magnifica-
tion). From this Figure, we can see that at low temperature
of 80˚C, a block-like morphology is found. As the proc-
ess is maturated, a morphological change from “blocky”
crystals to more needle-like crystals occurs, as shown in
Figure 3(b). Figures 3(a1, b1 ) shows high magnification
images of the as-grown materials. It is seen that hydro-
xyapatite precipitated uniformly on the surface to cover
the surface entirely and consists of needle-like crystals
radiating from a point in the form of a flower. Higher
temperature produces a higher crystallinity product. The
full width of one flower-like structure is about 50-60 m.
In addition, the HA nanorods are randomly grown and
they originate from the center of the flower. It seems that
the central part of the flower-shaped structures provides a
root for the growth of these HA nanorods. The precipita-
tion behavior of hydroxyapatite on the titanium plate at
low temperature resulted in block-like crystals and works
as a root for the growth of HA nanorods at higher tem-
perature.
Different morphology of crystal may imply the differ-
ent mechanism of formation. Several authors have sug-
gested that the electrolytes of lower OH concentration
usually accompanied with lower pH value and the octa-
calcium phosphate (OCP, Ca8H2(PO4)6·5H2O) could be a
precursor phase for HAp formation. Brown and Smith [24]
suggested that OCP is the original precipitate on which
Copyright © 2011 SciRes. MSA
Effect of Hydroxide Ion Concentration on the Morphology of the Hydroxyapatite Nanorods Synthesized Using 109
Electrophoretic Deposition
biological apatite nucleates. In the present work, the elec-
trolyte associated with ribbon-like HAp has a pH value of
4.0, more acidic than the electrolyte depositing rod-like
HAp (pH = 10). Therefore, it might be possible that in the
electrolyte of lower OH concentration the OCP was
formed during the deposition and took the morphology of
needle-like.
The XRD patterns of the representative HAP samples
are shown in Figure 4. This Figure shows that the two
coatings exhibit typical apatite peaks at 2θ of 25.9˚ and
31-33˚, which correspond to the HAp (002) diffraction
and a combination of the poorly resolved (211), (202) and
(300) diffractions. The peaks from both coatings are in
good agreement with the standard HAp patterns. All peaks
could be indexed to a hexagonal HAp crystal. No char-
acteristic peaks of impurities, such as calcium hydroxide
and calcium phosphates, were observed, meaning that
phase pure HAp was prepared under the present experi-
mental conditions.
FT-IR studies also confirmed that the coatings consist
of HAp crystals (Figure 5). The stretching mode of the
OH group appears at 3572 cm1. The bands that corre-
spond to the internal modes of PO43 group occur at v1
(600, 570 cm1) and v2 (1085, 1033 cm1). The OH stretch-
ing band at 3572 cm1 is considered as the indication of
HAp structure [25].
4. Conclusions
A simple and controllable synthesis, without surfactants,
of hexagonal hydroxyapatite nanorods by using electro-
phoretic deposition has been reported. The hydroxyapa-
tite nanorods were formed on a titanium electrode at
temperatures ranging from 80 to 130˚C.
1) The morphology of hydroxyapatite crystal can be
effectively controlled by the concentration of OH in the
electrolytes prepared for the deposition.
2) When OH concentration was 104, a highly uniform
nanorod HAp crystal can be deposited on the titanium
surface. The size of the deposited nanorods increased
remarkably with increasing electrolyte temperature, While,
in the electrolyte with lower OH concentration of 1010, a
needle-like flower-shaped HA can be prepared.
3) In the electrolyte with lower OH concentration of
1010, a block-like HA crystals were obtained at electro-
lyte temperatures of 80˚C, while a gradual change to
needle-like flower-shaped HA occurred at 100˚C elec-
trolyte temperature.
4) The high resolution FE-SEM images and selected
area electron diffraction patterns showed that the nanos-
tructures obtained are single crystalline with hexagonal
structure, grown along the direction of the c-axis.
Figure 4. The XRD patterns of the representative HAp na-
norods. (a) HA Flowershape; (b) HA Nanorods.
Figure 5. The FTIR spectra of HA synthesized under dif-
ferent OH concentration.
5. Acknowledgements
This work was supported by Research Center, College of
Engineering, King Saud University, Kingdom of Saudi
Arabia under Project No. 23/430.
REFERENCES
[1] Y. M. Kong, S. Kim and H. E. Kim, “Reinforcement of
Hydroxyapatite Bioceramic by Addition of ZrO2 Coated
with Al2O3,” Journal of the American Ceramic Society,
Vol. 82, No. 11, 1999, pp. 2963-2968.
doi:10.1111/j.1151-2916.1999.tb02189.x
[2] W. Suchanek and M. Yoshimura, “Processing and Prop-
erties of Hydroxyapatite-Based Biomaterials for Use as
Hard Tissue Replacement Implants,” Journal of Materials
Research, Vol. 13, No. 1, 1998, pp. 94-117.
doi:10.1557/JMR.1998.0015
[3] M. H. Fathi and A. Hanifi, “Evaluation and Characteriza-
Copyright © 2011 SciRes. MSA
Effect of Hydroxide Ion Concentration on the Morphology of the Hydroxyapatite Nanorods Synthesized Using
Electrophoretic Deposition
Copyright © 2011 SciRes. MSA
110
28
tion of Nanostructure Hydroxyapatite Powder Prepared
by Simple Sol-Gel Method,” Materials Letters, Vol. 61,
No. 18, 2007, pp. 3978-3983.
doi:10.1016/j.matlet.2007.01.0
kamoto and K. Ioku[4] M. Yoshimura, H. Suda, K. O,
“Hydrothermal Synthesis of Biocompatible Whiskers,”
Journal of Materials Science, Vol. 29, No. 1994, pp.
3399-3402.
doi:10.1007/BF00352039
Synthesis of Inorganic Materi-
a0
[5] S. Mann and G. A. Ozin, “
als with Complex Form,” Nature, Vol. 382, No. 6589,
1996, pp. 313-318.
doi:10.1038/382313
kana and P. Yang, “Langmuir-
151-2916.1998.tb02529.x
[6] F. Kim, S. Kwan, J. A
Blodgett, Nanorod Assembly,” Journal of the American
Ceramic Society, Vol. 123, No. 18, 2001, pp. 4360-4361.
[7] Y. Ota, T. Iwashit, T. Kasuga and Y. Abe, “Novel Prepa-
ration Method of Hydroxyapatite Fibers,” Journal of the
American Ceramic Society, Vol. 81, No. 6, 1998, pp.
1665-1668.
doi:10.1111/j.1
rotein Chroma-
/0003-9861(56)90183-7
[8] A. Tiselius, S. Hjertén and O. Levin, “P
tography on Calcium Phosphate Columns,” Archives of
Biochemistry and Biophysics, Vol. 65, No. 1, 1956, pp.
132-155.
doi:10.1016
tites. II. Preparation [9] O. Fowler, “Infrared Studies of Apa
of Normal and Isotopically Substituted Calcium, Stron-
tium, and Barium Hydroxyapatites and Spectra-Structure-
Composition Correlations,” Inorganic Chemistry, Vol. 13,
No. 1, 1974, pp. 207-214.
doi:10.1021/ic50131a040
[10] A. Bigi, E. Boanini and K. Rubini, “Hydroxyapatite Gels
c.2004.05.018
and Nanocrystals Prepared through a Sol-Gel Process,”
Journal of Solid State Chemistry, Vol. 177, No. 9, 2004,
pp. 3092-3098.
doi:10.1016/j.jss
Synthesis [11] M. Li, H. Schnablegger and S. Mann, “Coupled
and Self-Assembly of Nanoparticles to Give Structures
with Controlled Organization,” Nature, Vol. 402, No.
6760, 1999, pp. 393-395.
doi:10.1038/46509
[12] D. Walsh, J. D. Hopwood and S. Mann, “Construction of
1576
Reticulated Calcium Phosphate Frameworks in Bicon-
tinuous Reverse Microemulsions,” Science, Vol. 264, No.
5165, 1994, pp. 1576-1578.
doi:10.1126/science.264.5165.
g and G. Wang, “A [13] Y. Liu, W. Wang, Y. Zhan, C. Zhen
Simple Route to Hydroxyapatite Nanofibers,” Materials
Letters, Vol. 56, No. 4, 2002, pp. 496-501.
doi:10.1016/S0167-577X(02)00539-6
[14] S. Bose and S. K. Saha, “Synthesis and Characteriz
21/cm0303437
ation
of Hydroxyapatite Nanopowders by Emulsion Technique,”
Chemistry of Materials, Vol. 15, No. 23, 2003, pp. 4464-
4469.
doi:10.10
ann, “Synthesis of Barium Sul-[15] J. D. Hopwood and S. M
fate Nanoparticles and Nanofilaments in Reverse Micelles
and Microemulsions,” Chemistry of Materials, Vol. 9, No.
8, 1997, pp. 1819-1828.
doi:10.1021/cm970113q
[16] X. Peng, L. Manna, W. Yang, J. Wickham, E. Scher, A.
38/35003535
Kadavanich and A. P. Alivisatos, “Shape Control of CdSe
Nanocrystals,” Nature, Vol. 404, No. 6773, 2000, pp.
59-61.
doi:10.10
eart, S. O. Obare, C. J. Johnson, K. [17] N. R. Jana, L. A. Gearh
J. Edler, S. Mann and C. J. Murphy, “Liquid Crystalline
Assemblies of Ordered Gold Nanorods,” Journal of Ma-
terials Chemistry, Vol. 12, No. 10, 2002, pp. 2909-2912.
doi:10.1039/b205225c
[18] Z. Liu, Z. Hu, J. Liang, S. Li, Y. Yang, S. Peng and Y.
/la035160d
Qian, “Size-Controlled Synthesis and Growth Mechanism
of Monodisperse Tellurium Nanorods by a Surfactant-As-
sisted Method,” Langmuir, Vol. 20, No. 1, 2004, pp.
214-218.
doi:10.1021
and A. S. Posner, “Crystal Struc-
050a0
[19] M. I. Kay, R. A. Young
ture of Hydroxyapatite,” Nature, Vol. 204, No. 4963, 1964,
pp. 1050-1052.
doi:10.1038/2041
The Devitrification Behavior of [20] J.-S. Lee and C.-K. Hsu, “
Calcium Phosphate Glass with TiO2 Addition,” Thermo-
chimica Acta, Vol. 333, No. 2, 1999, pp. 115-119.
doi:10.1016/S0040-6031(99)00095-7
[21] W. J. Li, E. W. Shi, W. Z. Zhong and Z. W. Yin, “Growth
Mechanism and Growth Habit of Oxide Crystals,” Journal
of Crystal Growth, Vol. 203, No. 1-2, 1999, pp. 186-196.
doi:10.1016/S0022-0248(99)00076-7
[22] G. C. Koumoulidis, T. C. Vaimakis and A. T. Sdoikos,
“Preparation of Hydroxyapatite Lathlike Particles Using
High-Speed Dispersing Equipment,” Journal of the Ameri-
can Ceramic Society, Vol. 84, No. 6, 2001, pp. 1203-1208.
doi:10.1111/j.1151-2916.2001.tb00817.x
[23] J. N. Nian and H. Teng, “Hydrothermal Synthesis of Sin-
gle-Crystalline Anatase TiO2 Nanorods with Nanotubes
as the Precursor,” The Journal of Physical Chemistry B,
Vol. 110, No. 9, 2006, pp. 4193-4198.
doi:10.1021/jp0567321
[24] W. E. Brown and J. P. Smith, “Octacalcium Phosphate
and Hydroxyapatite: Crystallographic and Chemical Relations
between Octacalcium Phosphate and Hydroxyapatite,” Na-
ture, Vol. 196, 1962, pp. 1050-1055.
doi:10.1038/1961050a0
[25] M. Ashok, N. M. Sundaram and S. N. Kalkura., “Crystal-
lization of Hydroxyapatite at Physiological Temperature,”
Materials Letters, Vol. 57, No. 13-14, 2003, pp. 2066-2070.
doi:10.1016/S0167-577X(02)01140-0