You are on page 1of 11

Ceramics International 44 (2018) 17612–17622

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Silicon-doped hydroxyapatite prepared by a thermal technique for hard T


tissue engineering applications
Maryam Tavafoghia, Joseph M. Kinsellab, Cé Guinto Gamysc, Mathilde Gosselinc,

Yaoyao Fiona Zhaoa,
a
Department of Mechanical Engineering, McGill University, Montreal, QC, Canada
b
Department of Bioengineering, McGill University, Montreal, QC, Canada
c
Materium Innovations Company, Granby, QC, Canada

A R T I C LE I N FO A B S T R A C T

Keywords: Hydroxyapatite (HA, Ca5(PO4)3OH) has been extensively used for bone implantation due to its similarity to the
Si-doped hydroxyapatite mineral component of bone, which makes it strongly osteoconductive. However, HA has low resorbability, and it
Resorbability is difficult to replace by a newly regenerated bone. Si doping can enhance the resorbability of HA by modifying
Mineralization its crystal structure. Here, we developed a simple thermal technique for preparing Si-doped HA from silica (SiO2)
Hard tissue engineering
and HA precursors, both of which are inexpensive and commercially available. This method included the
physical binding of SiO2 and HA particles, followed by pressing and sintering the mixture at an elevated tem-
perature, which enhanced the atomic diffusion of Si into HA unit cells. We also evaluated the simulated body
fluid (SBF) activity of the Si-doped HA prepared by this technique and showed that it significantly had higher
resorbability and mineralizing potential compared to the pure HA. Our experimental design including, the in-
dividual precipitation and resorption assays enabled us to explain the mechanism behind the improved activity
of Si-doped HA in SBF. This was attributed to the formation of new phases, such as β-tricalcium phosphate (β-
TCP) and calcium silicate (Ca2SiO4) with higher solubility than HA on the SiO2-contating HA during the sintering
stage. This can provide some guidelines for designing new calcium phosphate-based materials for hard tissue
engineering applications.

1. Introduction other hand, is shown to enhance the resorbability of HA by modifying


its morphology and crystal structure [8–10]. The Si-doped HA is also
Hydroxyapatite (HA, Ca5(PO4)3OH) is the main inorganic compo- shown to have enhanced in vitro [10,11] and in vivo [8,12] bone re-
nent of bone and teeth. Early studies on the use of HA in hard tissue generation properties and cell proliferation capability [13,14] com-
engineering go back to the 1950s, and since that time, the synthesis of pared to the stoichiometric HA. In addition, the presence of silicon itself
HA as a potential material for bone implantation and replacement has can promote bone formation and calcification, and it is generally noted
been extensively investigated [1–4]. Synthetic HA has been extensively that silicon is essential for the growth and development of biological
used for bone grafting due to its similarity to the inorganic component tissues, such as bone and teeth [15]. Overall, an ideal set of properties,
of bone, which makes it extremely osteoconductive [5,6]. As opposed to such as bioactivity and biocompatibility combined with an acceptable
bioglasses especially borate glasses, HA is highly biocompatible, and it range of mechanical strength can make Si-doped HA a more competent
possesses higher mechanical strength compared to many other bioac- osteogenic agent compared to bioactive glasses and biopolymers [8].
tive glasses and some types of calcium phosphates. This can make HA a Aqueous precipitation techniques [15–19] are commonly used for
better candidate for load bearing sites [6]. However, the low resorb- synthesizing Si-doped HA. However, these chemical techniques are
ability of HA compared to the bioglasses and its difficulty to replace by complex in terms of optimizing the processing conditions, and usually
a newly generated bone in vivo has been increasingly limited the have poor ability in controlling the crystallinity of the produced Si-
clinical use of HA in bone scaffolds [7,8]. In fact, clinical investigations doped HA particles. Also, they are not efficient for producing Si-doped
have shown that HA implants are virtually inert and can remain in body HA on a large scale. The Si-doped HA can be alternatively produced by
for approximately six years after implantation [6]. Si doping, on the thermal techniques [20–23], where HA and a Si-containing precursor


Corresponding author.
E-mail address: yaoyao.zhao@mcgill.ca (Y.F. Zhao).

https://doi.org/10.1016/j.ceramint.2018.06.071
Received 29 March 2018; Received in revised form 9 June 2018; Accepted 9 June 2018
Available online 15 June 2018
0272-8842/ © 2018 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

are mixed physically and sintered at temperatures high enough to en- Table 1
hance the atomic diffusion of Si into HA. However, the bone re- The list of samples and their preparation conditions.
generation properties of the Si-doped HA material produced by such Name Component (theoretical values) Processing
thermal techniques has been hardly investigated.
In this work, we investigated the in vitro resorbability and miner- HA HA As received powder
HA-E HA HA immersed in
alizing potential of Si-doped HA produced by a simple thermal tech-
ethanol
nique. This method included the physical mixing of HA and silica (SiO2) HA-G HA HA-E after grinding
powder by the aid of an organic binder, followed by pressing and sin- SiO2 SiO2 As received powder
tering the mixture at an elevated temperature, which enhanced the SiO2-E SiO2 SiO2 immersed in
atomic diffusion of Si into HA unit cells. The sintered compacts were ethanol
SiO2-G SiO2 SiO2-E after grinding
then characterized by a variety of techniques to verify the formation of
HA-GP HA (97 wt%)+ PEG (3 wt%) Green powder
Si-doped HA on these samples. Finally, the compacts were immersed in HA-GC Green compact
simulated body fluid (SBF) to investigate their dissolution and miner- HA-CC HA (100 wt%) + SiO2 (0 wt%) Calcined compact
alization behavior at physiological mimicking conditions. Differently HA-SC Si: 0 wt%, Si: 0 at% Sintered compact
HA-(2% SiO2)- HA (1.9 wt%) + SiO2 ( 95.1 wt%) + Green powder
from other studies on Si-doped HA materials, here we provided a pre-
GP PEG (3 wt%), HA:SiO2 = 98:2 wt%,
cise experimental design, including the individual mineralization and HA-(2% SiO2)- Green compact
dissolution assays. This enabled us to explain in more detail the me- GC
chanism behind the improved bone regeneration activity of Si-doped HA-(2% SiO2)- HA (98 wt%) + SiO2 (2 wt%) Calcined compact
HA. Our goal was to develop a simple and cost-effective technique for CC
HA-(2% SiO2)- Si: 0.94 wt%, Si: 0.77 at% Sintered compact
preparing HA-based materials with improved bone regeneration prop-
SC
erties for hard tissue engineering applications. SiO2-GP SiO2 (97 wt%)+ PEG (3 wt%) Green powder
SiO2-GC Green compact
2. Materials and method SiO2-CC HA (0 wt%) + SiO2 (100 wt%) Calcined compact
SiO2-SC Si: 46.76 wt%, Si: 33.33 at% Sintered compact

2.1. Materials

HA powder (Ca5(PO4)3OH, purity ≥ 95%) was purchased from 400 °C for 5 h in air to remove the binder and then sintered in an
Berkeley Advanced Biomaterials Inc., and amorphous silica micro- electrical furnace (ExOne) at 1100 °C for 8 h in air at a heating and
spheres (SiO2, purity > 95%) were received from Materium cooling rate of 10 °C/min. The weight of compacts was measured by a
Innovations (MATSPHERE™MAT300). More details about Materium digital analytical balance, and the dimensions were measured by a di-
company can be found in the affiliation. Materium also assisted in gital caliper. Absolute (ρ) and relative (ρr) density of compacts were
characterizing the initial SiO2 and HA powder. The HA and silica par- calculated according to the following equations for at least 3 samples
ticle sizes were 50–150 and 5–40 µm, and their surface areas were 10 for each set, and the values were average.
and > 100 m2/g, respectively. Polyethylene glycol 8000 (PEG, purity
ρ = m/v (1)
≥ 95%) and ethanol (≥ 99.5%) were purchased from Fisher Scientific.
ρr = ρ/ρth (2)
2.2. Preparing SiO2-containing HA powder
ρth = (ρth of SiO2) × (wt% of SiO2) + (ρth of HA) × (wt% of HA) (3)
Adequate amounts of as-received silica (3 g) and HA powder (147 g)
with SiO2:HA weight ratio of 2:98 were immersed in 250 ml of ethanol. where, m and v are mass and volume of compacts, respectively; ρth is
4.6 g (resulting in 3 wt% of total weight) of a binder, PEG, was dis- theoretical density of compacts, and ρth of HA and SiO2 is 3.156 and
solved in 30 ml of deionized (DI) water and then added to the HA and 2.648 g/cm3, respectively.
silica mixture to bind the silica and HA particles together. The same
procedure was followed for 150 g of as-received SiO2 and HA powders 2.4. Mineralization and dissolution experiments
as control samples, and the binder was also added to keep consistency
with the SiO2-containing HA powder. The solutions were then kept SBF was prepared according to a well-known recipe provided by
stirring on a magnetic stirrer with the rate of 500 rpm until the ethanol Kokubo et al. [24]. Also, to evaluate HA dissolution, a Ca- and P-defi-
evaporated partially and then they were dried in the oven at 60 °C cient SBF solution (so called DSBF) was prepared by eliminating Ca-
overnight. The dried mixtures were ground manually by a mortar and and P-containing precursors from SBF. It should be noted that Ca
pestle to obtain green powders (GPs), HA-GP, HA-(2% SiO2)-GP, and (2.5 mM) and P (1 mM) concentration in SBF is very low [24], and the
SiO2-GP as shown in Table 1. To investigate the changes in the mor- elimination of these ions would not have a significant impact on the
phology of powders during this process, the same procedure was fol- ionic strength (0.16 vs 0.15 for SBF and DSBF, respectively) and ki-
lowed for as-received HA and SiO2 powder without adding any binder. netics of calcium phosphate dissolution. The ionic strength in body fluid
These powders were called SiO2-EtOH and HA-EtOH after immersion, is mainly regulated by NaCl, significant amount of which (137.5 mM)
and SiO2-G and HA-G after grinding. was present in both SBF and DSBF. Also, both SBF and DSBF contained
a buffer reagent (tris, 50.5 mM), which kept the pH constant at about
2.3. Si-doped HA preparation 7.4 during the course of reaction. The sintered compacts prepared by
pellet-pressing (SC samples) were immersed in 40 ml of SBF or DSBF
Cylindrical compacts (15 mm in diameter, and 6–7 mm in thickness solution in centrifugation tubes and the tubes were placed in an in-
depending on the loose density of initial powders) were prepared by cubator at 37 °C for 15 days. The solutions were agitated gently every 2
uniaxial pressing at 5 metric ton in a steel die using 2 g of each HA-GP, days. The samples were finally removed from the incubator, washed 3
HA-(2% SiO2)-GP and SiO2-GP powder. The compacts were calcined at times with DI water, and dried in the oven at 60 °C.

17613
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

Fig. 1. BSE-SEM (b and c) and SEM (a and d–l) images of HA-(2% SiO2)-GP (a-c), HA (d and e), HA-E (f), HA-G (g), HA-GP (h), SiO2 (i),SiO2-E (j), SiO2-G (k), and
SiO2-GP (l) samples. The red and blue arrows on panel b and c show the HA and SiO2 particles, respectively, that were identified by EDS analysis. (For interpretation
of the references to colour in this figure legend, the reader is referred to the web version of this article.)

2.5. Characterizations The dilution with 4% nitric acid was to make sure that no more pre-
cipitate formed in the solutions after filtration. The Ca, P, and Si con-
2.5.1. Inductively coupled plasma atomic emission spectroscopy (ICP-AES) centrations were measured at the wavelength of 317.9, 178.3, and
Ca, P, and Si concentration in the SBF and DSBF solutions con- 212.4 nm, respectively. Dissolved amounts of Ca (Cad) and P (Pd) at
taining HA-SC, HA-(2% SiO2)-SC, and SiO2-SC samples were measured different incubation times were measured from DSBF solution that did
at different incubation times using an ICP-AES instrument (ICAP 6500 not contain any Ca and P-containing precursors. To keep consistency,
Duo). 1 ml aliquots were taken from the SBF and DSBF solutions at the the dissolved amounts of Si (Sid) were also measured from DSBF solu-
desired incubation times, filtered with a syringe filter with a pore size of tion. For this, the initial amounts of Si (SiiDSBF), P (PiDSBF), and Ca
100 nm to remove any trace of precipitates or particles released from (CaiDSBF) in blank DBSF were subtracted from the measured amounts of
the samples, and then immediately diluted 5 times with 4% nitric acid. these elements at different incubation times in DSBF. The precipitated

17614
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

amounts of Ca (Cap) and P (Pp) were also calculated according to the mean. Statistical analyses were performed by Student's T-Test with
following equations, p < 0.05 considered significant.
Cap (mol) = CaiSBF + Ca d − CaSBF (4)
3. Results and discussions
Pp (mol) = PiSBF + Pd − PSBF (5)
Si was doped into HA by sintering, and the effect of Si doping on the
where CaiSBF and PiSBF are the initial amounts of Ca and P in blank SBF, sinterability and SBF activity of HA was studied. A wide range of SiO2
and CaSBF and PSBF are the measured amounts of Ca and P in SBF so- concentrations from 0 to 100 wt% were investigated, and it was shown
lution at different incubation times. that 2 wt% SiO2 was the optimum concentration, where the SBF activity
Also, the dissolution and mineralization of compacts after 14 day in terms of the changes observed in Ca and P concentration was sig-
immersion in DSBF and SBF, respectively, were investigated by calcu- nificantly different from 0 wt% SiO2 (Fig. S1), while the mechanical
lating the compacts weight changes according to the following equa- properties of compacts were not yet deteriorated (Fig. S2). Therefore,
tions, this concentration of SiO2 was chosen for the rest of the analyses.
(mDSBF − m s)
Dissolution(%) = × 100 3.1. The homogeneity of SiO2-containing HA powder
mS (6)

mSBF + (ms − mDSBF) − m s SiO2-containing HA powder (HA-(2% SiO2)-GP) was prepared ac-
Mineralization(%) = × 100
mS (7) cording to a technique explained in Section 2.2. SEM images of HA-(2%
SiO2)-GP are shown in Fig. 1a–c. The dark gray particles observed more
where ms is the mass of sintered compacts (SCs), and mDSBF, and mSBF
clearly on the BSE-SEM images (see blue arrows on Fig. 1b and c) were
are the mass of SCs after 14 day immersion in DSBF and SBF, respec-
identified by EDS (3 spectra at least) to be SiO2 and the lighter particles
tively.
(see red arrows on Fig. 1c) were HA. The quantitative EDS analysis was
performed on the circular area of 25 µm in diameter, and the results
2.5.2. Fourier transform infrared (FT-IR) spectroscopy indicated that within this range, Si was homogeneously dispersed in HA
IR spectra were recorded on a Bruker Tensor 27 FT-IR spectrometer at a concentration of 0.8 ± 0.1 at%, which was very close to theore-
using diffuse reflectance (DRIFT) mode. The samples were diluted with tical concentration of Si at 0.77 at%.
KBr to approximately 10 wt/wt%. Pure KBr powder was used as a Overall, the particles observed on the green powders, HA-(2%
background. The FT-IR spectra were recorded from 600 to 4000 cm−1 SiO2)-GP (Fig. 1a–c), HA-GP (Fig. 1h), and SiO2-GP (Fig. 1l), were
using a deuterated triglycine sulfate (DTGS) detector. The spectra were much smaller (~0.5–12 µm) than those observed on the as-received
collected by averaging 256 scans at 4 cm−1 resolution. SiO2 (~25 µm) (Fig. 1i) and HA (~100 µm) (Fig. 1d and e) powders. To
investigate the changes in the morphology and size of particles during
2.5.3. X-ray diffraction (XRD) the preparation of green powders, the pure (no binder was added) HA
XRD spectra were collected using a Bruker AXS XRD instrument and SiO2 powders seperately underwent the same procedure as the
with Cu Kα radiation generated at 40 kV and 40 mA. A quartz and green powders. The results indicated that only immersing and stirring
metallic sample holder was used for the powders and compacts, re- the HA (Fig. 1f) and SiO2 (Fig. 1j) powders in ethanol was enough to
spectively. Diffraction angles collected (2θ) ranged from 5° to 54°, with break down the particles and make them smaller, while no significant
step size of 0.05°. changes were observed in the ground HA (Fig. 1g) and SiO2 (Fig. 1k)
compared to the ethanol-immersed powders. The presence of large
2.5.4. X-ray photoelectron spectroscopy (XPS) aggregates on these samples could be attributed to electrostatic forces
XPS measurements were performed using a monochromatic X-ray usually observed between small particles.
photoelectron spectrometer K Alpha (Thermo Scientific). The setup was
equipped with an Al Kα X-Ray source (1486.6 eV, 0.834 nm), a micro- 3.2. Effect of SiO2 addition on the morphology and sinterability of HA
focused monochromator and an ultrahigh vacuum chamber (10−9 compacts
Torr). The survey scans were collected with energy steps of 1 eV and
spot size of 400 µm. Scans were taken on at least 3 points on each The green compacts (GC samples) were made by pressing the HA-
sample and the quantitative results were averaged. A flood gun was GP, HA-(2% SiO2)-GP, and SiO2-GP powders listed in Table 1. These
used to neutralize electrical built-up charge generated. Peak fitting and compacts were then calcined (CC samples) to remove the organic
quantitative analysis of survey spectra were performed using the soft- binder and sintered (SC samples) at 1100 °C for 8 h to make the com-
ware Thermo Avantage (version 4.60). pacts stronger. In this section, we show the changes in the morphology
and composition of compacts during these thermal processes, and also
2.5.5. Scanning electron microscopy (SEM) investigate the effect of SiO2 addition on the sinterability of the HA
The morphology of samples was analyzed with a scanning electron compacts. Sintering at elevated temperatures can also make the Si
microscope (SEM) from Hitachi (S-4700 FE-SEM) using variable pres- atoms become incorporated into the HA lattice [10,17]. This will be
sure (VP) mode and an acceleration voltage of 10 kV. Energy dispersive discussed in detail in Section 3.3.
spectrometer (EDS) was used for the elemental analyses. EDS spectra The SEM images of CC and SC samples containing 0, 2, and 100 wt%
were collected on at least 5 points on each sample and the quantitative of SiO2 are shown in Fig. 2. While the morphology of particles in the HA
results were averaged. The specimens were mounted on a double sided compact did not change considerably during sintering (Fig. 2a and b),
conductive carbon tape and were not metal coated. the SiO2 compact underwent significant changes during the sintering
stage (Fig. 2e and f), where closely packed particles and well defined
2.6. Statistical analysis grain boundaries were observed on SiO2-SC (Fig. 2f). This indicated a
higher sinterability of SiO2 compared to HA powder, which is also
All quantitative results are shown as standard deviation of the confirmed by the density results shown in Fig. 3: both the absolute

17615
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

Fig. 2. SEM images of HA-CC (a), HA-SC (b), HA-SC immersed in SBF (c), HA-SC immersed in DSBF (d), SiO2-CC (e), SiO2-SC (f), SiO2-SC immersed in SBF (g), SiO2-
SC immersed in DSBF (h), HA-(2% SiO2)-CC (i), HA-(2% SiO2)-SC (j), HA-(2% SiO2)-SC immersed in SBF (k), and HA-(2% SiO2)-SC immersed in DSBF (l). The blue
arrows on panel i show the SiO2 particles that were identified by EDS analysis. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

(Fig. 3a) and relative (Fig. 3b) density of SiO2 compact increased investigated in more detail by SEM-EDS (Fig. 4a) and XPS (Fig. 4b)
drastically after sintering while the density of the HA compact did not compositional analyses that were performed on the compacts surfaces.
change significantly. Also, adding 2 wt% SiO2 to HA did not help im- While Si content in HA-(2% SiO2)-GP was estimated to be
prove the sinterability of this powder (Fig. 3a and b): both the absolute 0.8 ± 0.1 at% by EDS, the Si concentration within the same analysis
(2.05 vs 2.16 g/cm3) and relative (65.3 vs 68.4%) density measured on range was higher at 2.0 ± 0.2 and 2.8 ± 0.3 at% on the surface of HA-
HA-(2% SiO2)-SC was lower than on HA-SC. This could be attributed to (2% SiO2)-GC and HA-(2% SiO2)-CC, respectively. This indicated the
the low concentration of SiO2, or to the fact that SiO2 was not homo- homogeneity of SiO2 within the HA-(2% SiO2)-GP was disrupted during
genously dispersed in HA. In fact, SiO2 aggregates of ~ 15 µm in size the compaction and calcination stage and it became more concentrated
were observed on HA-(2% SiO2)-CC (Fig. 2i); however, these aggregates on the surface likely due to the fact that SiO2 has a lower density than
disappeared after sintering on HA-(2% SiO2)-SC (Fig. 2j). This was HA (2.65 vs 3.16 g/cm3). However, the Si content on HA-(2% SiO2)

17616
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

decreased from 2.8 ± 0.3 to 1.8 ± 0.1 at% after sintering (Fig. 4a), HA-(2% SiO2)-CC (Fig. 5a, iii) showed almost all characteristic
which indicated that sintering at an elevated temperature (1100 °C) peaks relative to its HA component and it also showed some major
could make the SiO2 migrate from the surface into the compacts. To peaks of amorphous silica at 812, 1105, and 1215 cm−1. However,
confirm this, XPS analysis, which is more sensitive to surface compo- upon sintering (Fig. 5a, iv), the peak at 812 cm−1 shifted to 795 cm−1
sition than EDS, was performed. As expected, the Si concentration on HA-(2% SiO2)-SC due to the formation of crystalline silica. Also, the
measured by XPS (Fig. 4b) on HA-(2% SiO2)-CC was significantly peak at 1215 cm−1 corresponding to ν Si-O-Si [36] in silica dis-
higher than the one measured by EDS (Fig. 4a). However, consistent appeared, and the peak at 962 cm−1 originating from ν1 PO4 [29] in HA
with the EDS results, XPS results also confirmed that the Si measured on became weaker. Instead, some new peaks emerged at 498 and
HA-(2% SiO2)-SC (4.2 ± 0.1 at%) was lower than on HA-(2% SiO2)-CC 947 cm−1 on HA-(2% SiO2)-SC, which could be attributed to δ SiO4 and
(10.1 ± 0.6 at%) indicating the effect of sintering on improving the Si ν SiO4, respectively [34,38–40]. These IR results indicated the decrease
homogeneity. These compositional analyses indicated that the homo- in the SiO2 content of HA-(2%SiO2)-SC sample due to the Si diffusion
geneity of Si distribution on HA-(2% SiO2)-GP decreased during the into HA during sintering and the formation of SiO44- groups, which
compaction process; however, sintering helped to improve the homo- substituted for PO43- in HA. This could result in the formation of Si-
geneity of the compacts, but Si was still more concentrated on the doped HA (Ca10(PO4)6-x(SiO4)x(OH)2-x) according to Eq. (8)
surface than inside the compacts. [10,17,38,39]. The lower concentration of Si on the surface of HA-(2%
Overall, the results reported in the literature indicate that Si addi- SiO2)-SC compared to HA-(2% SiO2)-CC (Fig. 4) can further confirm
tion has an adverse effect on the sinterability of HA scaffolds this hypothesis. SiO44- observed on HA-(2%SiO2)-SC could also origi-
[18,25,26]. Our results as well confirmed this observation; however, nate from calcium silicate (Ca2SiO4) that can form following the de-
the deteriorating effect of 2 wt% SiO2 on sinterability of HA was neg- composition of Si-doped HA (Eq. (9)) [10]. However, IR results did not
ligible here (Fig. 3). The densities obtained for HA-SC (2.16 ± 0.03 g/ evidently show the formation of TCP, which is also expected to form
cm3) and HA-(2% SiO2)-SC (2.05 ± 0.02 g/cm3) samples were both according to Eq. (9). To further investigate the formation of new
significantly higher than the mineral density of trabecular bone crystalline phases during sintering, XRD analysis (Fig. 5b) was per-
(~ 0.2 g/cm3) [27], and were more comparable to that of cortical bone formed.
(~ 2.0 g/cm3) [28]. Yet, the relative densities obtained for HA-SC Consistent with the IR results, the XRD spectra for the HA sample
(68.4 ± 0.8%) and HA-(2% SiO2)-SC (65.3 ± 0.8%) were still lower did not change significantly during sintering (Fig. 5b, i and ii), while
than the maximum density of ~ 98% reported for HA-based compacts the initially amorphous SiO2 became completely crystalline (Fig. 5b, v
prepared under the same conditions [15,18]. This can be attributed to and vi). This was evident from the sharp peaks on SiO2-SC, which
the low sinterability of our HA-GP powder, which had large particle matched well the XRD reference pattern for crystalline silica. HA-(2%
sizes and diverse morphologies. SiO2)-CC (Fig. 5b, iii) only showed the peaks relative to HA because
SiO2 was amorphous at this stage; however, HA-(2% SiO2)-SC (Fig. 5b,
3.3. Si-doping into HA by sintering iv) showed both the HA and SiO2 peaks since SiO2 became crystalline
during sintering. HA-(2% SiO2)-SC also showed some relatively strong
The SiO2 and HA powders were mixed physically, so only the peaks corresponding to beta-tricalcium phosphate (β-TCP) and calcium
characteristic features related to these two materials were expected to silicate as expected to form according to Eq. (9) following the decom-
be present on the HA-(2% SiO2) samples. However, sintering the sam- position of Si-doped HA [10,17]. In fact, the XRD peaks of HA are ex-
ples at an elevated temperature could enhance the atomic diffusion of Si pected to shift to lower angles upon the incorporation of Si into the HA
into the HA lattice, thus resulting in the formation of new crystalline lattice [41]. However, since such changes were not evident on our high
phases according to the following reactions [10,17], resolution XRD spectra (Fig. S3), it can be deduced that the Si-doped
HA was totally decomposed into TCP and calcium silicate during the
Ca10 (PO4)6 (OH)2 +xSiO44 − → Ca10 (PO4)6 − x (SiO4 ) x (OH)2 − x + xPO34− relatively long sintering time (8 h).
+xOH− (8)
3.4. Effect of SiO2 addition on SBF activity of HA
x
Ca10 (PO4)6 − x (SiO4 ) x (OH)2 − x → ⎛1− ⎞ Ca10 (PO4 )6 (OH)2 +xCa3 (PO4 )2
⎝ 2⎠ Fig. 6a–g shows the amount of Ca, P, and Si that are dissolved from
the compacts in DSBF (Fig. 6a–d), or are removed from SBF due to the
+xCa2SiO4 (9)
calcium phosphate precipitation on the compacts (Fig. 6e–g), and
The formation of new phases on the samples during sintering was Fig. 6h shows the Ca/P ratio in the precipitates. These values are
investigated by IR (Fig. 5a) and XRD (Fig. 5b) techniques. Table 2 measured by ICP according to a technique that is discussed in detail in
shows the assignment of peaks found on FTIR spectra recorded on HA- Section 2.5.1. The total dissolution and mineralization were also in-
CC, HA-SC, HA-(2% SiO2)-CC, HA-(2% SiO2)-SC, SiO2-CC, and SiO2-SC vestigated by measuring the changes in the weight of compacts after
samples [29–36]. The HA sample (Fig. 5a, i and ii) did not undergo immersion in DSBF or SBF, respectively (Fig. 6i).
significant changes during sintering; only a small peak at 716 cm−1
corresponding to OH [35] appeared on HA-SC, which could be attrib- 3.4.1. HA dissolution
uted to the increase in crystallinity HA by sintering [37]. However, HA-(2% SiO2)-SC showed significantly higher Ca dissolution than
SiO2-SC (Fig. 5a, vi) showed more significant changes compared to HA-SC that did not contain any silica (Fig. 6a). However, P dissolution
SiO2-CC (Fig. 5a, v): overall, the peaks became sharper on SiO2-SC. on HA-(2% SiO2)-SC was less than on HA-SC (Fig. 6b), but it should be
Also, the δ Si-O-Si band at 812 cm−1 on SiO2-C shifted to 795 cm−1 on considered that P dissolution for all samples was almost negligible
SiO2-SC, and a small peak at 623 cm−1 corresponding to δ Si-O-Si ap- (< 1 µmol). HA-(2% SiO2)-SC also showed a significantly higher Si
peared on SiO2-SC [33]. The δ Si-O-Si bands at 623 and 795 cm−1 dissolution compared to HA-SC (Fig. 6c). The Si dissolution on HA-(2%
observed on SiO2-SC are the features of silicate chains ((SiO3)∞) [34] SiO2)-SC was even faster than the pure silica samples, SiO2-SC. Overall,
and can be attributed to the formation of crystalline silica after sin- HA-(2% SiO2)-SC showed the fastest dissolution rate in DSBF, followed
tering. by HA-SC and SiO2-SC (Fig. 6d and i). The faster dissolution rate of HA-

17617
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

(2% SiO2)-SC was also evident from the SEM images (Fig. 2). While the
surface morphology of HA-SC (Fig. 2d) and SiO2-SC (Fig. 2h) immersed
in DSBF was not significantly different from HA-SC (Fig. 2b) and SiO2-
SC (Fig. 2f), the surface of HA-(2% SiO2)-SC (Fig. 2j) became rougher
and pitted after immersion in DSBF (Fig. 2l).

3.4.2. Calcium phosphate precipitation


Precipitation results in Fig. 6e–g and i indicated that the calcium
phosphate precipitate formed on all samples; however, precipitation on
SiO2-SC was almost negligible. This was also evident from the SEM
image of the SiO2-SC immersed in SBF (Fig. 2g). The rate of calcium
phosphate precipitation on HA-(2% SiO2)-SC was significantly higher
than on HA-SC; however, the particles of the newly formed calcium
phosphate were not distinguishable on the SEM images of either HA-
(2% SiO2)-SC (Fig. 2k) or HA-SC (Fig. 2c) immersed in SBF. This is
while, spherulitic HA particles are usually expected to form on biolo-
gical surfaces immersed in SBF [11,17,30]. This could be attributed to
the fact that the precipitates that formed on the samples were removed
during the washing stage, or possibly the precipitating particles here
had a similar morphology to those already present on the compacts,
thus making them undistinguishable. However, the XPS results in
Fig. 4b showed that Si concentration on HA-(2% SiO2)-SC immersed in
SBF was significantly lower than on both the HA-(2% SiO2)-SC, and the
HA-(2% SiO2)-SC immersed in DSBF. This could be attributed to the
presence of a layer of newly formed calcium phosphate on HA-(2%
SiO2)-SC immersed in SBF [42]. Therefore, the latter scenario seems to
be more likely to happen considering the fact that the formation of
similar particles to those on HA-SC and HA-(2% SiO2)-SC is also re-
Fig. 3. Absolute (a) and relative (b) density of HA, HA-(2% SiO2), and SiO2
ported by other authors on the Si-doped HA scaffolds placed in SBF
compacts after calcination (CC samples) and sintering (SC samples). Asterisk [10,11,42,43].
signs (*) show the values that were statically significantly different with
P < 0.05. 3.4.3. How Si doping can improve HA resorbability and mineralization?
It is shown that Si doping, even in a small amount at ~ 0.8 wt%, can
significantly improve the SBF activity of HA [10,11,16,38,42,44].
Consistent with these results, we showed that our Si-doped HA, pro-
duced by sintering the 2 wt% SiO2-containng HA compacts, had higher
SBF activities than the pure HA (Fig. 6). Such SBF activity tests are
widely used for predicting the bioactivity of biomaterials in vivo [16].
However, SBF studies on Si-doped HA have rarely been able to distin-
guish between the mineralizing and dissolving phases [10,11,42].
Moreover, these studies merely reported the changes in [Ca] and [P] in
SBF rather than measuring the amounts of precipitating and dissolving
Ca and P [10,11,38,42]. Introducing a Ca- and P-deficient solution (so
called DSBF) enabled us to investigate the dissolution, which was fol-
lowed by some simple calculations (explained in detail in Section 2.5.1)
to measure the amounts of dissolving and precipitating Ca and P. This
allowed us to better understand the mechanism of calcium phosphate
mineralization or dissolution in the presence of Si-doped HA in SBF. For
example, the higher resorbability of Si-doped HA is usually attributed to
the partial transformation of HA to TCP with a higher resorbability than
HA [10]. However, our results showed that while Ca dissolution was
significantly higher for HA-(2% SiO2)-SC than for HA-SC, the P dis-
solution for both of these samples was comparable and almost negli-
gible. This indicated that the higher solubility of HA-(2% SiO2)-SC
cannot be attributed to the faster dissolution of TCP formed on this
sample during sintering (Fig. 5b). On the other hand, XRD results
(Fig. 5b) showed that other phases, such as Ca2SiO4 were also present
on HA-(2% SiO2)-SC. Ca2SiO4 (5.8 ×10−4 M at 20 °C) has a higher
solubility than both HA (3.2 ×10−5 M at 37 °C) and TCP (1.6 ×10−4 M
at 37 °C) [45], and considering the significantly higher amounts of Ca
and Si dissolution for HA-(2% SiO2)-SC than for HA-SC (Fig. 5a and c),
it can be concluded that the dissolution of Ca2SiO4 was mainly re-
Fig. 4. Si at% measured on HA, HA-(2% SiO2), and SiO2 samples after calci- sponsible for the higher resorbability of HA-(2% SiO2)-SC. Also, the Ca/
nation (CC samples), sintering (SC samples), and immersion in SBF and DSBF by Si ratio of 1.6 ± 0.3 measured based on the dissolved Ca and P (Fig. 6a
either SEM-EDS (a) or XPS (b). Asterisk signs (*) show the values that were and c) for HA-(2% SiO2)-SC after 14 day incubation was close to that
statically significantly different with P < 0.05.
for Ca2SiO4 (2), which further confirms the dissolution of Ca2SiO4.

17618
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

Fig. 5. FTIR spectra (a) and XRD pat-


terns (b) recorded on HA-CC (i), HA-SC
(ii), HA-(2% SiO2)-CC (iii), HA-(2%
SiO2)-SC (iv), SiO2-CC (v), and SiO2-SC
(vi) samples. The numbers in black and
blue on panel a indicate the peaks ori-
ginating from HA and SiO2, respectively,
and the numbers in red indicate new
phases that were formed during the
sintering stage and are discussed in de-
tail in Section 3.3. The XRD patterns are
compared with reference spectra for HA
(black lines, PDF card n. 00-009-0432),
SiO2 (blue lines, PDF card n. 00-041-
0539), β-TCP (red lines, database code
AMCSD 0007447), and β-Ca2SiO4 (green
lines, database code AMCSD 0020214).
(For interpretation of the references to
colour in this figure legend, the reader is
referred to the web version of this ar-
ticle.)

Table 2 glasses can act as intermediating crystals for the formation of bone like
Assignment of peaks found on FTIR spectra shown in Fig. 5 for HA-CC, HA-SC, crystals [47–49], and this too can be responsible for the faster miner-
HA-(2% SiO2)-CC, HA-(2% SiO2)-SC, SiO2-CC, and SiO2-SC. alization of HA-(2% SiO2)-SC sample. In fact, Si is an essential trace
Peaks Wavenumbers (cm−1) element in many organs of animals, and its presence plays a significant
role in bone mineralization through promoting collagen type 1 synth-
HA-CC HA-SC HA-(2% HA-(2% SiO2-CC SiO2-SC esis and enhancing osteoblast differentiation [6,16,50,51]; therefore, Si
SiO2)-CC SiO2)-SC dissolution can also play a role in improving HA mineralization at in
ν2 PO4 432 474 432 474 – – vivo conditions.
474 474 To identify the precipitating calcium phosphate, the Ca/P ratios
δ Si-O-Si – – – – 476 476 (Fig. 6h) were measured for the precipitates based on the amount of
δ SiO4 – – – 498 – – precipitating Ca and P shown in Fig. 6e and f. The Ca/P for HA-SC was
ν4 PO4 575 575 575 575 – –
almost constant at 1.8 ± 0.1 during the whole incubation period in
604 604 604 604
635 635 635 635 SBF. This Ca/P ratio can be attributed to the formation of carbonated
OH – 716 – 716 – – HA (CHA), which has Ca/P ratios ranging from 1.67 to 2 depending on
δ Si-O-Si – – 812 – 812 – its carbonate content [52]. The Ca/P for HA-(2% SiO2)-SC was
δ Si-O-Si in – – – 795 – 795
2.0 ± 0.1 after 3 day incubation and it increased to 2.2 ± 0.1 after 15
(SiO3)∞ 623
ν1 CO3 891 891 891 891 – – days, which could be attributed to the incorporation of more carbonate
ν SiO4 – – – 947 – – ions into the HA precipitate forming on this sample. The Ca/P was
ν1 PO4 962 962 962 962 – – significantly high for SiO2-SC at 3 day incubation (8.2 ± 0.7) and it
ν3 PO4 1028 1028 1028 1028 – – decreased to 2.7 ± 0.2 after 15 days. The high Ca/P ratio here could be
1057 1057 1057 1057
attributed to the formation of Ca(OH)2 derivatives [53] on SiO2-SC;
1092 1092 1092 1092
ν Si-O-Si – – 1105 1105 1105 1105 however, it should be considered that the precipitation on SiO2-SC was
1215 1215 1215 almost negligible (Figs. 2g and 6g).
To further evaluate the precipitating material, XRD analysis was
performed on the samples after immersion in SBF (Fig. S5). No calcium
However, the XRD patterns of HA-(2% SiO2)-SC immersed in SBF or phosphate phases were detected on the XRD pattern of SiO2-SC im-
DSBF (Fig. S4) did not show any decline in the peaks related to Ca2SiO4 mersed in SBF, which indicates the precipitation on SiO2-SC was not
due to the fact that the main peaks of Ca2SiO4 overlaps with HA. significant. Also, the XRD patterns of HA-SC and HA-(2% SiO2)-SC did
Total calcium phosphate precipitation also was higher for HA-(2% not show significant changes after immersion in SBF. This can indicate,
SiO2)-SC than for HA-SC (Fig. 6g). This can be explained by the higher the newly formed calcium phosphates (as were already confirmed by
dissolution rate of these samples, which could increase the local con- ICP analyses in Fig. 6f and g) were probably HA or TCP, which were
centration of Ca in SBF and favor calcium phosphate mineralization on already present on the samples. However, the precipitation of TCP from
the samples [11,46]. Also, it is shown that the silane groups in bioactive SBF is not favored at physiological conditions [24]. Also, the Ca/P

17619
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

Fig. 6. Dissolution and mineralization of HA-SC, HA-(2% SiO2)-SC, and SiO2-SC scaffolds immersed in DSBF and SBF, respectively, at different incubation times
evaluated by ICP: (a, b, c, d) Ca (Cad), P (Pd), Si (Sid), and total (Cad + Pd + Sid) dissolution of scaffolds in DSBF; (e, f, g) Ca (Cap), P (Pp), and total calcium phosphate
(Cap + Pp) precipitation from SBF. (h) Ca/P ratio in the precipitated phase. (i) scaffold weight changes after 14 day immersion in DSBF or SBF corresponding to
dissolution or mineralization, respectively. Asterisk (*) and plus (+) signs indicate the values that are statically significantly different from the correspondent values
measured on HA-SC and SiO2-SC, respectively with P < 0.05.

ratios (Fig. 6h) measured for HA-SC (~ 1.8) and HA-(2% SiO2)-SC conditions [16,57]. Also, the resorbability of HA-(2% SiO2)-SC was still
(2–2.2) were closer to that for CHA (1.67–2) than for TCP (1.5). This significantly lower than some bioglasses such as borosilicate and borate
can indicate the formation of HA on these samples, which is also con- bioactive glasses that showed dissolution rate of 0.06 and 0.3 wt% h−1,
firmed by other researches to form on HA-based scaffolds immersed in respectively, in SBF [58]. However, it should be noted that the low
SBF [54,55]. degree of porosity in our Si-doped HA samples produced by a com-
Overall, HA-(2% SiO2)-SC showed significantly improved both the paction technique could also be responsible for its lower activity at
resorbability and mineralizing potential compared to the pure HA in biological conditions.
physiological conditions (Fig. 6), and this would make the Si-doped HA
a more ideal candidate than HA for hard tissue engineering applica- 4. Conclusion
tions, where both the scaffold resorbability and regenerability is re-
quired. The dissolution and mineralization rate of ~ 0.001 wt% h−1 In this work, we shed some light on the mechanism under which Si
obtained for HA-(2% SiO2)-SC sample was comparable to the values doping can improve the resorbability and mineralizing potential of HA.
reported for the Si-doped HA produced by aqueous precipitation tech- This can provide some guidelines for designing new HA-based materials
niques and tested in vitro [56]; however, they were approximately 30 to be used for hard tissue engineering applications.
times lower than when the bioactivity tests were conducted in in vivo We showed the Si-doped HA prepared by sintering 2 wt% SiO2-

17620
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

contating HA powder had both the higher mineralization and dissolu- characterization of silicon-doped hydroxyapatite, Mater. Trans. 50 (5) (2009)
tion rate in SBF compared to the pure HA. This is while the deterior- 1046–1049.
[18] I.R. Gibson, S.M. Best, W. Bonfield, Effect of silicon substitution on the sintering and
ating effect of 2 wt% SiO2 on sinterability and mechanical properties of microstructure of hydroxyapatite, J. Am. Ceram. Soc. 85 (11) (2002) 2771–2777.
HA compacts was almost negligible. The enhanced SBF activities of 2 wt [19] I.R. Gibson, S.M. Best, W. Bonfield, Chemical characterization of silicon-substituted
% SiO2-contating HA were attributed to the formation of Si-doped HA hydroxyapatite (in eng), J. Biomed. Mater. Res. 44 (4) (1999) 422–428.
[20] S. Langstaff, M. Sayer, T.J.N. Smith, S.M. Pugh, S.A.M. Hesp, W.T. Thompson,
on this sample at elevated sintering temperatures. The decomposition of "Resorbable bioceramics based on stabilized calcium phosphates. Part I: rational
Si-doped HA consequently led to the formation of new phases, such as design, sample preparation and material characterization, Biomaterials 20 (18)
β-TCP and Ca2SiO4, which had higher water solubility than HA. Also, (1999) 1727–1741.
[21] J.W. Reid, A. Pietak, M. Sayer, D. Dunfield, T.J.N. Smith, Phase formation and
the presence of silane groups on Si-doped HA could be responsible for evolution in the silicon substituted tricalcium phosphate/apatite system,
the faster mineralization of this sample. Biomaterials 26 (16) (2005) 2887–2897.
[22] M. Sayer, et al., Structure and composition of silicon-stabilized tricalcium phos-
phate, Biomaterials 24 (3) (2003) 369–382.
Acknowledgment
[23] X.W. Li, H.Y. Yasuda, Y. Umakoshi, Bioactive ceramic composites sintered from
hydroxyapatite and silica at 1200 °C: preparation, microstructures and in vitro
This research project is supported by the Natural Science and bone-like layer growth (in eng), J. Mater. Sci. Mater. Med. 17 (6) (2006) 573–581.
Engineering Research Council of Canada (NSERC) EGP 469410-2014. [24] T. Kokubo, H. Takadama, How useful is SBF in predicting in vivo bone bioactivity?
Biomaterials 27 (15) (2006) 2907–2915.
Also, we would like to thank Prof. Marta Cerruti from Materials [25] A. Bianco, I. Cacciotti, M. Lombardi, L. Montanaro, Si-substituted hydroxyapatite
Engineering, McGill University for providing access to a FT-IR instru- nanopowders: Synthesis, thermal stability and sinterability, Mater. Res. Bull. 44 (2)
ment. (2009) 345–354.
[26] S.R. Kim, et al., Synthesis of Si, Mg substituted hydroxyapatites and their sintering
behaviors, Biomaterials 24 (8) (2003) 1389–1398.
Appendix A. Supplementary material [27] Y. Jiang, et al., Trabecular bone mineral and calculated structure of human bone
specimens scanned by peripheral quantitative computed tomography: relation to
biomechanical properties (in eng), J. Bone Miner. Res 13 (11) (1998) 1783–1790.
Supplementary data associated with this article can be found in the [28] D.D. Thompson, Age changes in bone mineralization, cortical thickness, and
online version at http://dx.doi.org/10.1016/j.ceramint.2018.06.071. Haversian canal area, Calcif. Tissue Int. 31 (1) (1980) 5–11.
[29] E. Garskaite, K.-A. Gross, S.-W. Yang, T.C.-K. Yang, J.-C. Yang, A. Kareiva, Effect of
processing conditions on the crystallinity and structure of carbonated calcium hy-
References droxyapatite (CHAp), CrystEngComm 16 (19) (2014) 3950–3959, http://dx.doi.
org/10.1039/C4CE00119B.
[1] A. Perloff, A.S. Posner, Preparation of pure hydroxyapatite crystals, Science 124 [30] M. Tavafoghi, N. Brodusch, R. Gauvin, M. Cerruti, Hydroxyapatite formation on
(3222) (1956) 583–584. graphene oxide modified with amino acids: arginine versus glutamic acid, J. R. Soc.
[2] M.I. Kay, R.A. Young, A.S. Posner, Crystal structure of hydroxyapatite, Nature 204 Interface 13 (114) (2016), http://dx.doi.org/10.1098/rsif.2015.0986.
(4963) (1964) 1050–1052, http://dx.doi.org/10.1038/2041050a0. [31] S. Roushdey, Defect Related Luminescence in Silicon Dioxide Network: A Review,
[3] M.T. Jahromi, G. Yao, M. Cerruti, The importance of amino acid interactions in the 2011.
crystallization of hydroxyapatite, J. R. Soc. Interface 10 (80) (2012), http://dx.doi. [32] G. Dingemans, C.A.A. Van Helvoirt, D. Pierreux, W. Keuning, W.M.M. Kessels,
org/10.1098/rsif.2012.0906. Plasma-assisted ALD for the conformal deposition of SiO2: process, material and
[4] M. Tavafoghi Jahromi, M. Cerruti, Amino acid/ion aggregate formation and their electronic properties, J. Electrochem. Soc. 159 (3) (2012) H277–H285.
role in hydroxyapatite precipitation, Cryst. Growth Des. 15 (3) (2015) 1096–1104. [33] J.F.B. Barata, A.L. Daniel-Da-Silva, M.G.P.M.S. Neves, J.A.S. Cavaleiro, T. Trindade,
[5] J.R. Woodard, et al., The mechanical properties and osteoconductivity of hydro- Corrole-silica hybrid particles: synthesis and effects on singlet oxygen generation,
xyapatite bone scaffolds with multi-scale porosity, Biomaterials 28 (1) (2007) RSC Adv. 3 (1) (2013) 274–280, http://dx.doi.org/10.1039/C2RA22133K.
45–54. [34] P. Yu, R.J. Kirkpatrick, B. Poe, P.F. Mcmillan, X. Cong, Structure of calcium silicate
[6] G. Kaur, O.P. Pandey, K. Singh, D. Homa, B. Scott, G. Pickrell, A review of bioactive hydrate (C-S-H): near-, mid-, and far-infrared spectroscopy, J. Am. Ceram. Soc. 82
glasses: their structure, properties, fabrication and apatite formation (in Eng), J. (3) (1999) 742–748.
Biomed. Mater. Res A 102 (1) (2014) 254–274. [35] L. Berzina-Cimdina, N. Borodajenko, Research of Calcium Phosphates Using Fourier
[7] M.B. Conz, J.M. Granjeiro, G. d.A. Soares, Hydroxyapatite crystallinity does not Transform Infrared Spectroscopy, INTECH Open Access Publisher, Croatia, 2012.
affect the repair of critical size bone defects, J. Appl. Oral. Sci. 19 (4) (2011) [36] G.F. Andrade, D.C.F. Soares, R.K.D.S. Almeida, E. Sousa, S.M. Barros, Mesoporous
337–342 (Jul-Aug 03/22/received 11/20/revised 10/26/accepted). silica SBA-16 functionalized with Alkoxysilane groups: Preparation, characteriza-
[8] A.E. Porter, N. Patel, J.N. Skepper, S.M. Best, W. Bonfield, Comparison of in vivo tion, and release profile study, J. Nanomater. 2012 (2012) 10 (Art. no. 816496.).
dissolution processes in hydroxyapatite and silicon-substituted hydroxyapatite [37] P.E. Wang, T.K. Chaki, Sintering behaviour and mechanical properties of hydro-
bioceramics (in eng), Biomaterials 24 (25) (2003) 4609–4620. xyapatite and dicalcium phosphate, J. Mater. Sci.: Mater. Med., vol. 4(2), pp.
[9] A.E. Porter, C.M. Botelho, M.A. Lopes, J.D. Santos, S.M. Best, W. Bonfield, 150–158.
Ultrastructural comparison of dissolution and apatite precipitation on hydro- [38] L.T. Bang, B.D. Long, R. Othman, Carbonate hydroxyapatite and silicon-substituted
xyapatite and silicon-substituted hydroxyapatite in vitro and in vivo (in eng), J. carbonate hydroxyapatite: synthesis, mechanical properties, and solubility evalua-
Biomed. Mater. Res A 69 (4) (2004) 670–679. tions, Sci. World J. 2014 (2014) 9 (Art. no. 969876).
[10] Y. Belmamouni, M. Bricha, J. Ferreira, K. El Mabrouk, Hydrothermal synthesis of Si- [39] D. Arcos, J. Rodrı́guez-Carvajal, M. Vallet-Regı, The effect of the silicon in-
doped hydroxyapatite nanopowders: mechanical and bioactivity evaluation, Int. J. corporation on the hydroxylapatite structure. A neutron diffraction study, Solid
Appl. Ceram. Technol. 12 (2) (2015) 329–340. State Sci. 6 (9) (2004) 987–994.
[11] E.S. Thian, J. Huang, S.M. Best, Z.H. Barber, W. Bonfield, Novel silicon-doped hy- [40] P. Saravanapavan, L.L. Hench, Mesoporous calcium silicate glasses. I. Synthesis, J.
droxyapatite (Si-HA) for biomedical coatings: an in vitro study using a cellular si- Non-Cryst. Solids 318 (1–2) (4/1/ 2003) 1–13.
mulated body fluid (in eng), J. Biomed. Mater. Res B Appl. Biomater. 76 (2) (2006) [41] A. El Yacoubi, A. Massit, M. Fathi, B.C. El Idrissi, K. Yamni, Characterization of
326–333. silicon-substituted hydroxyapatite powders synthesized by a wet precipitation
[12] N. Patel, et al., A comparative study on the in vivo behavior of hydroxyapatite and method, IOSR J. Appl. Chem. 7 (11) (2014) 24–29.
silicon substituted hydroxyapatite granules (in eng), J. Mater. Sci. Mater. Med. 13 [42] F. Balas, J. Perez-Pariente, M. Vallet-Regi, In vitro bioactivity of silicon-substituted
(12) (2002) 1199–1206. hydroxyapatites (in eng), J. Biomed. Mater. Res A 66 (2) (2003) 364–375.
[13] J.L. Xu, K.A. Khor, Chemical analysis of silica doped hydroxyapatite biomaterials [43] M.S. Sadjadi, H.R. Ebrahimi, M. Meskinfam, K. Zare, Silica enhanced formation of
consolidated by a spark plasma sintering method, J. Inorg. Biochem. 101 (2) (2007) hydroxyapatite nanocrystals in simulated body fluid (SBF) at 37 °C, Mater. Chem.
187–195. Phys. 130 (1–2) (2011) 67–71.
[14] M. Palard, J. Combes, E. Champion, S. Foucaud, A. Rattner, D. Bernache-Assollant, [44] G. Munir, G. Koller, L. Di Silvio, M.J. Edirisinghe, W. Bonfield, J. Huang, The
Effect of silicon content on the sintering and biological behaviour of Ca10(PO4)(6- pathway to intelligent implants: osteoblast response to nano silicon-doped hydro-
x)(SiO4)x(OH)(2-x) ceramics (in eng), Acta Biomater. 5 (4) (2009) 1223–1232. xyapatite patterning, J. R. Soc. Interface 8 (58) (2011) 678–688, http://dx.doi.org/
[15] S.R. Kim, et al., Synthesis of Si, Mg substituted hydroxyapatites and their sintering 10.1098/rsif.2010.0548.
behaviors (in eng), Biomaterials 24 (8) (2003) 1389–1398. [45] J. Rakovan, Growth and surface properties of apatite, Rev. Mineral. Geochem. 48
[16] Z.-Y. Qiu, I.-S. Noh, S.-M. Zhang, Silicate-doped hydroxyapatite and its promotive (1) (2002) 51–86.
effect on bone mineralization, Front. Mater. Sci. 7 (1) (2013) 40–50. [46] J. Weng, Q. Liu, J.G. Wolke, X. Zhang, K. De Groot, Formation and characteristics of
[17] K. Nakata, T. Kubo, C. Numako, T. Onoki, A. Nakahira, Synthesis and the apatite layer on plasma-sprayed hydroxyapatite coatings in simulated body
fluid (in eng), Biomaterials 18 (15) (1997) 1027–1035.

17621
M. Tavafoghi et al. Ceramics International 44 (2018) 17612–17622

[47] S. Hayakawa, K. Tsuru, C. Ohtsuki, A. Osaka, Mechanism of apatite formation on a [53] S. Raynaud, E. Champion, D. Bernache-Assollant, P. Thomas, Calcium phosphate
sodium silicate glass in a simulated body fluid, J. Am. Ceram. Soc. 82 (8) (1999) apatites with variable Ca/P atomic ratio I. Synthesis, characterisation and thermal
2155–2160. stability of powders (in eng), Biomaterials 23 (4) (2002) 1065–1072.
[48] M. Vallet-Regí, A.M. Romero, C.V. Ragel, R.Z. LeGeros, XRD, SEM-EDS, and FTIR [54] H.M. Kim, T. Himeno, M. Kawashita, T. Kokubo, T. Nakamura, The mechanism of
studies of in vitro growth of an apatite-like layer on sol-gel glasses, J. Biomed. biomineralization of bone-like apatite on synthetic hydroxyapatite: an in vitro as-
Mater. Res. 44 (4) (1999) 416–421. sessment, J. R. Soc. Interface 1 (1) (2004) 17–22.
[49] C.A. Miller, T. Kokubo, I.M. Reaney, P.V. Hatton, P.F. James, Formation of apatite [55] T. Kokubo, H. Takadama, How useful is SBF in predicting in vivo bone bioactivity?
layers on modified canasite glass–ceramics in simulated body fluid, J. Biomed. Biomater. Rev. 27 (15) (2006) 2907–2915.
Mater. Res. 59 (3) (2002) 473–480. [56] S. Sprio, et al., Physico-chemical properties and solubility behaviour of multi-sub-
[50] E.M. Carlisle, Silicon: a possible factor in bone calcification (in eng), Science 167 stituted hydroxyapatite powders containing silicon, Mater. Sci. Eng.: C 28 (1) (1/
(3916) (1970) 279–280. 10/ 2008) 179–187.
[51] D.M. Reffitt, et al., "Orthosilicic acid stimulates collagen type 1 synthesis and os- [57] A.M. Pietak, J.W. Reid, M.J. Stott, M. Sayer, Silicon substitution in the calcium
teoblastic differentiation in human osteoblast-like cells in vitro (in eng), Bone 32 (2) phosphate bioceramics, Biomaterials 28 (28) (2007) 4023–4032.
(2003) 127–135. [58] Q. Fu, M.N. Rahaman, H. Fu, X. Liu, Silicate, borosilicate, and borate bioactive glass
[52] J.P. Lafon, E. Champion, D. Bernache-Assollant, Processing of AB-type carbonated scaffolds with controllable degradation rate for bone tissue engineering applica-
hydroxyapatite Ca10−x(PO4)6−x(CO3)x(OH)2−x−2y(CO3)y ceramics with con- tions. I. Preparation and in vitro degradation, J. Biomed. Mater. Res. Part A 95A (1)
trolled composition, J. Eur. Ceram. Soc. 28 (1) (2008) 139–147. (2010) 164–171.

17622

You might also like