Microwave Assisted Synthesis and Characterization of Magnesium Substituted Calcium Phosphate Bioceramics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science and Engineering C 56 (2015) 286–293

Contents lists available at ScienceDirect

Materials Science and Engineering C

journal homepage: www.elsevier.com/locate/msec

Microwave assisted synthesis and characterization of magnesium


substituted calcium phosphate bioceramics
Nida Iqbal Khan a,b, Kashif Ijaz a, Muniza Zahid a, Abdul S. Khan a, Mohammed Rafiq Abdul Kadir b,
Rafaqat Hussain c, Anis-ur-Rehman d, Jawwad A. Darr e, Ihtesham-ur-Rehman f, Aqif A. Chaudhry a,⁎
a
Interdisciplinary Research Centre in Biomedical Materials, COMSATS Institute of Information Technology, M. A. Jinnah Campus, Defence Road, Off Raiwind Road, Lahore, Pakistan
b
Medical Implant Technology Group (MEDITEG), Faculty of Bioscience and Medical Engineering, Universiti Teknologi Malaysia, 81310 Skudai, Johor Darul Takzim, Malaysia
c
Department of Chemistry, Faculty of Science, Universiti Teknologi Malaysia, Skudai, Johore, Malaysia
d
Department of Physics, COMSATS Institute of Information Technology, Chakshahzad Campus, Islamabad, Pakistan
e
Clean Materials Technology Group, Department of Chemistry, University College London, Christopher Ingold Laboratories, 20 Gordon Street, London WC1H 0AJ, UK
f
The Kroto Research Institute, North Campus, University of Sheffield, Broad Lane, Sheffield S3 7HQ, UK

a r t i c l e i n f o a b s t r a c t

Article history: Hydroxyapatite is used extensively in hard tissue repair due to its biocompatibility and similarity to biological ap-
Received 2 May 2014 atite, the mineral component of bone. It differs subtly in composition from biological apatite which contains other
Received in revised form 23 March 2015 ions such as magnesium, zinc, carbonate and silicon (believed to play biological roles). Traditional methods of hy-
Accepted 7 May 2015
droxyapatite synthesis are time consuming and require strict reaction parameter control. This paper outlines syn-
Available online 9 May 2015
thesis of magnesium substituted hydroxyapatite using simple microwave irradiation of precipitated suspensions.
Keywords:
Microwave irradiation resulted in a drastic decrease in ageing times of amorphous apatitic phases. Time taken to
Magnesium substitution synthesize hydroxyapatite (which remained stable upon heat treatment at 900 °C for 1 h) reduced twelve folds
Microwave synthesis (to 2 h) as compared to traditionally required times. The effects of increasing magnesium concentration in the
Bioceramics precursors on particle size, surface area, phase-purity, agglomeration and thermal stability, were observed
Calcium phosphates using scanning electron microscopy, BET surface area analysis, X-ray diffraction and photo acoustic Fourier trans-
form infra-red spectroscopy. Porous agglomerates were obtained after a brief heat-treatment (1 h) at 900 °C.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction stimulate osteoblast proliferation [20,21]. During synthesis magnesium


ions also bind to the surface of apatite crystals inhibiting their growth
Hydroxyapatite [HA] is a key bioceramic due to its excellent biocom- resulting in smaller particles (higher bioactivity due to higher surface
patibility and its bioactivity which stems from its compositional and area) [22,23]. Decrease in content of magnesium in bone and loss of
structural similarity to the mineral part of bone, enamel and dentine some crystallinity are associated with osteoporosis [24,25]. These
[1–4]. Although bone mineral is an apatitic phase it also contains addi- adverse effects related to magnesium deficiency can impair bone
tional ions which include magnesium, carbonate, strontium, zinc and growth and reduce bone quality (strength and density) and increase
silicate ions [5–7]. The type and amount of ionic substitution in biolog- bone fragility [5,20].
ical apatite varies from wt% level (e.g. CO23 −) to the ppm–ppb level As the magnesium ion is smaller than the a calcium ion (0.86 Å vs.
(Mg2+ or Sr2+) [8–10]. These ionic substitutions affect the crystal lat- 1.14 Å), the incorporation of magnesium affects the crystal structure
tice structure of HA and influence the dissolution rate and hence the of HA and substitution cannot occur over the full compositional range
resulting bioactivity of the bone mineral [11–13]. (i.e. magnesium ions cannot replace all calcium ions in HA) [26,27].
Magnesium substitution into HA is of interest due to its potential im- Substituting magnesium ions into HA lattice reduces crystal cell param-
pact on the mineralization process and crystallization of apatite [14–16]. eters and hence cell volume as expected; above a concentration limit of
It has been reported that in calcified tissues, the amount of magnesium ~0.15Mg2+/Ca2+ the apatite structure cannot accommodate any more
associated with the apatite phase is higher (5% at.) at the first stages of magnesium resulting in the formation of additional phases (typically
the bone remodeling process and decreases with increasing calcification TCP or amorphous calcium magnesium phosphates), that act as segre-
and ageing of an individual [17–19]. Moreover, magnesium can directly gating phases for this ion, is observed [28,29]. TCP phases are known
to have higher solubility than that of HA [30].
Several methods can be used for preparing calcium phosphates in-
⁎ Corresponding author. cluding magnesium substituted HA (Mg–HA), such as wet chemical
E-mail address: aqifanwar@ciitlahore.edu.pk (A.A. Chaudhry). routes based on precipitation at low temperature [22,31–36] or solid-

http://dx.doi.org/10.1016/j.msec.2015.05.025
0928-4931/© 2015 Elsevier B.V. All rights reserved.
N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293 287

state reactions at elevated temperatures [37,38] resulting in a range of chemical precipitation method for synthesis of HA can be expressed by
0.3–29 wt.% magnesium substitution. Traditional co-precipitation the following equation:
routes often require aging times of 18 h or more in order for the Ca:P
phases to reach maturation (and hence thermal stability) [39,40]. xMgðNO3 Þ2 þ ð10−xÞCaðNO3 Þ2 þ 6 ðNH4 Þ2 HPO4 →Ca10−x Mgx ðPO4 Þ6 ðOHÞ2
Other methods such as sol–gel syntheses [41], batch hydrothermal syn-
þ 12NH4 NO3 þ 8HNO3 :
theses [5], mechanochemical-hydrothermal synthesis methods [42,43]
and continuous hydrothermal flow synthesis [29] have been reported
to produce phase pure materials in the range of 0.5–28 wt.% magnesium 2.2. Synthesis methodology
substitution. Some of these techniques require expensive starting
materials, strict reaction parameter control and monitoring or require For a typical reaction, calcium and magnesium nitrate solutions were
expensive equipment. Therefore, a technique that reduces the time of prepared by dissolving appropriate amounts in 100 mL deionized water
hydroxyapatite synthesis without compromising on its desired proper- such that (Ca + Mg):P molar ratio was always ~1.67. A control sample
ties and also ensures sufficient ion substitution into hydroxyapatite (for structural comparison based on XRD and FTIR) labeled Ca–P was
would be of use. prepared by the same synthesis method but without any magnesium.
The advantages of microwave irradiation during chemical reactions Sample Ids and amounts of calcium nitrate and magnesium nitrate used
include acceleration of the reaction, smaller particle sizes with narrow are given in Table 1. Amounts were selected based on the following for-
particle distribution and finer nanostructures leading to improved me- mula for magnesium substituted HA [Mg–HA, Ca10 − xMgx (PO4)6OH2].
chanical properties when formed into dense test specimens [44–49]. Samples were denoted as follows: Ca–P corresponds to a sample without
Despite the benefits, ion substitutions using microwave assisted ap- any Mg (no magnesium nitrate added), 0.5Mg–CaP corresponds to
proaches have not been reported extensively in literature. Magnesium 0.5 wt.% Mg substituted in hydroxyapatite (based on theoretical 100%
[50] and silver [51] ion substitutions in HA have been reported using magnesium substitution into HA lattice). Similarly 2Mg–CaP corresponds
microwave assisted methods. The authors previously reported on the to a sample with 2 wt.% Mg substituted into HA. Therefore x and 10 − x
magnesium ion substitution into the HA lattice using Continuous represent Mg:Ca molar ratio. The sample labeled as Mg–P corresponds to
Hydrothermal Flow Synthesis (CHFS) Technology [29]. CHFS synthesis a powder made using magnesium nitrate and no calcium nitrate.
of Mg–HA exploited the benefits provided by near-critical water to re- 0.3 M diammonium hydrogen phosphate solution was prepared by
sult in rapid maturation (increase in Ca:P ratio from an initial amor- dissolving 3.96 g diammonium hydrogen phosphate in 100 mL deion-
phous precipitate). CHFS is a high throughput technique, however it ized water (pH adjusted to 10 by ammonium hydroxide addition).
requires an expensive setup which works at high temperature and pres- Diammonium hydrogen phosphate solution was added drop wise into
sure [52–54]. The authors have also previously reported on the rapid the calcium nitrate (if any) and magnesium nitrate (if any) solutions.
synthesis of thermally stable hydroxyapatite using microwave irradia- During addition, the pH of suspension (as instantaneous precipitation
tion and its deposition on metallic substrates [55,56]. In the current took place) was maintained at pH 10 by ammonium hydroxide addition
work, microwave irradiation was used as a much simpler alternative and once complete, the suspension was stirred for 30 min. The suspen-
method to controllably produce a range of magnesium substituted cal- sion was put into a modified household microwave oven (600 W,
cium phosphates with differing compositions and thermal stabilities 2.45 GHz) modified with a refluxing system. Microwave irradiation
in a rapid manner. was performed under ambient air for 5 min (15 s on and 15 s off)
[49]. The suspensions were observed to bubble at the base of the cooling
condenser (due to boiling by heating due to microwave irradiation).
2. Materials and methods After completion of exposure to microwave irradiation the suspension
was then filtered and washed with deionized water. The precipitate
2.1. Materials was dried at 80 °C in a drying oven for 24 h and sieved (below
74 μm). The sieved powder was then heat-treated at 900 °C for 1 h.
Diammonium hydrogen phosphate (NH4)2HPO4 (Applichem,
Darmstadt, Germany, 96–102%), calcium nitrate tetrahydrate Ca(NO3)2· 2.3. Characterization
4H2O (Unichem, Guangzzhou, China 99%), magnesium nitrate hexahy-
drate Mg(NO3)2·6H2O (Pamrea quimica SA, Barcelona, Spain 98%) and Powder X-ray diffraction (PXRD) data was collected on a X'Pert Pro
ammonium hydroxide NH4OH (Fisher Scientific, Loughborough, UK, PW3050/60 diffractometer in the 20–80° range, based on Cu K-α radia-
33%) were used. Deionized water was used in all reactions. The wet- tion [1.54060 Å], a step size of 0.02 and a scan time of 0.15 s per step.

Table 1
Sample IDs, corresponding x values, theoretical substitution weight percentages and amounts of reagents used.

Sample ID x (Mg) 10 − x (Ca) Mg:Ca molar ratio Mg wt % Added to 100 mL deionized water (grams)
(x:10 − x)
Calcium nitrate Magnesium nitrate

Concentration (M) Amount (grams) Concentration (M) Amount (grams)

CaP – 10 – – 0.500 11.800 – –


0.1Mg–CaP 0.04 9.96 0.004 0.1 0.497 11.700 0.002 0.054
0.5Mg–CaP 0.20 9.80 0.020 0.5 0.490 11.570 0.010 0.256
1Mg–CaP 0.40 9.60 0.042 1.0 0.480 11.330 0.020 0.513
1.5Mg–CaP 0.60 9.40 0.064 1.5 0.470 11.100 0.030 0.769
2Mg–CaP 0.80 9.20 0.087 2.0 0.460 10.860 0.040 1.026
4Mg–CaP 1.60 8.40 0.191 4.0 0.420 9.920 0.080 2.051
5Mg–CaP 2.00 8.00 0.250 5.0 0.400 9.447 0.100 2.564
10Mg–CaP 4.00 6.00 0.670 10.0 0.300 7.085 0.200 5.128
16Mg–CaP 6.00 4.00 1.500 16.0 0.200 4.724 0.300 7.692
22Mg–CaP 8.00 2.00 4.000 22.0 0.100 2.362 0.400 10.256
Mg–P 10.00 0.00 – 29.0 0.000 0.000 0.500 12.821
288 N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293

Table 2 3. Results and discussion


Magnesium substitution levels determined by area scans
using an EDX detector attached to a scanning electron
microscope.
3.1. Scanning electron microscopy

Sample ID Mg wt % age The chemical composition of as-precipitated samples was deter-


0.1Mg–CaP 0 mined using EDX and the results are shown in Table 2. The results indi-
0.5Mg–CaP 0.03 cate that the measured weight percentages of substituted magnesium
1Mg–CaP 0.97
were close to the theoretical amounts and followed an increasing trend.
1.5Mg–CaP 1.59
2Mg–CaP 1.64 Scanning electron microscopy was carried out to determine the size
4Mg–CaP 2.33 and morphology, degree of agglomeration and effect of heat-treatment.
5Mg–CaP 3.64 Fig. 1(a)–(d) shows the SEM images of heat-treated sample 1Mg–CaP
10Mg–CaP 7.3 at various magnifications. Particles were present in sintered agglomerat-
16Mg–CaP 11.86
22Mg–CaP 16.26
ed forms (and will hence be referred to as grains) with agglomerates as
Mg–P 22.83 large as ca. 90 μm as observed in Fig. 1(a) [with open porosity b 10 μm].
Fig. 1(b) and (c) revealed an average grain size of 618 nm [(±182 nm),
25 particles sampled]. Fig. 1(d) reveals these particles to be sintered po-
lygonal grains with distinct necking regions. This suggests that a certain
degree of agglomeration exists after precipitation; heat-treatment trans-
Fourier transform infra-red spectra were collected on a Thermo Nicolet forms these soft-agglomerates (statically bound particles) into harder
P6500 spectrometer using a photoacoustic cell purged with helium agglomerates (sintered particles). The literature suggests that indeed,
gas. 128 scans with 4 cm− 1 resolution were collected for each sam- granular (porous agglomerates) are a preferred form of calcium phos-
ple. Powder morphology and particle size was investigated using a phate ceramics for clinical use [57–59].
scanning electron microscope (JEOL JSM6490A, Japan). UTHSCSA A further increase in magnesium content from 1 wt.% (sample 1 Mg–
image tool was used to measure grain sizes from SEM images (grains CaP) to 5 wt.% (5Mg–CaP) resulted in formation of smaller agglomerates
with distinct grain boundaries were selected for sizing). Samples for with minimal porosity [compared to Fig. 1(a)] as seen in Fig. 2(a). The
SEM analysis were gold plated with coating thickness 250 Å using coat- SEM image in Fig. 2(b) revealed grains with an average size of 261 nm
ing unit JFC 1500 Ion Sputtering Device [JEOL ltd., Tokyo, Japan]. EDX [(±89 nm), 25 grains sampled]. Fig. 2(c)–(d) shows the SEM images
analysis was carried out using an EDX detector attached to a Hitachi for sample 10Mg–CaP revealing similar agglomeration and an average
scanning electron microscope SU-1500. BET Surface Area was deter- grain size of 256 nm [(± 58 nm), 25 grains sampled]. Samples 5Mg–
mined using N2 adsorption in a Micrometrics TriStar Surface Area Ana- CaP and 10Mg–CaP therefore have similar morphology and sizes. Ag-
lyzer. Samples were degassed at 100 °C under N2 purging for 2 h prior to glomerates of sample 10Mg–CaP were observed to be nanoporous
analysis. (pores b 0.5 μm).

Fig. 1. Scanning electron microscope images of heat treated (at 900 °C for 1 h) sample 1Mg–CaP at (a) ×1k magnification [bar = 10 μm] (b) at ×5k magnification [bar = 5 μm], (c) at ×10k
magnification [bar = 1 μm] and (d) at ×20k magnification [bar = 1 μm].
N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293 289

Fig. 2. Scanning electron microscope images of heat treated Mg–CaP samples 5 Mg–CaP at (a) ×5k magnification [bar = 5 μm] and (b) ×50k magnification [bar = 0.5 μm], 10Mg–CaP at
(c) ×5k magnification [bar = 5 μm] and (d) ×50k magnification [bar = 0.5 μm] and 22Mg–CaP at (e) ×5k magnification [bar = 5 μm] and (f) ×50k magnification [bar = 0.5 μm].

A further increase in nominal magnesium content from 10 to 22 wt.% samples. As explained later, an increase beyond 2 wt.% magnesium re-
(for sample 22Mg–CaP) resulted in a size dispersion of seemingly sulted in the disappearance of the broad peak characteristic of poorly
non-porous agglomerates (1 μm to 10 μm) as seen in Fig. 2(e)–(f). The crystalline apatitic phases in XRD patterns for as-precipitated samples
average grain size was measured to be 144 nm [(±32), 25 grains [seen in Fig. 4(a)–(f) but not observed in Fig. 4(g)–(l)]. This could be
sampled]. attributed to different phases with more amorphous nature being pre-
cipitated (hence the decrease in surface area). It is therefore suggested
that the decrease in agglomerate size and their dispersion [as seen in
3.2. BET surface area Fig. 2(e)–(f)] is possibly due to reduced sinterability as a result of
lower surface area.
Fig. 3 shows the BET surface area of as-precipitated and heat-treated
samples as a function of nominal magnesium weight percentages (in
the solution). The highest surface area (146 m2/g) for as-precipitated 3.3. Powder X-ray diffraction
samples was observed for sample 1.5Mg–CaP. Indeed, magnesium
has been reported to decrease the particle size of calcium phosphates Powder X-ray diffraction (PXRD) data was collected for all samples
[29]. In our case this holds true for samples 0.1Mg–CaP, 0.5Mg–CaP, for assessing phase purity and crystallinity. Fig. 4 shows the PXRD pat-
1Mg–CaP and 1.5Mg–CaP only. Further increase in nominal magne- terns of as-precipitated samples. PXRD patterns for samples Ca–P,
sium content resulted in a decreasing surface area for as-precipitated 0.1Mg–CaP, 0.5Mg–CaP, 1Mg–CaP, 1.5Mg–CaP and 2Mg–CaP in
290 N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293

Fig. 5. Powder X-ray diffraction patterns of heat treated (at 900 °C for 1 h) samples (a) Ca–P
(b) 0.1Mg–CaP (c) 0.5Mg–CaP (d) 1.0Mg–CaP (e) 1.5Mg–CaP (f) 2.0Mg–CaP (g) 4Mg–CaP
(h) 5Mg–CaP (i) 10Mg–CaP (j) 16Mg–CaP (k) 22Mg–CaP and (l) Mg–P (Note: HA = ,
Fig. 3. BET surface area of as-precipitated and heat treated (at 900 °C for 1 h) samples as a
Mg-βTCP = , Stanfieldite = , Farringtonite = ).
function of theoretical magnesium weight percentages (in solution).

treatment at 900 °C for 1 h [49]. A further increase in nominal magne-


Fig. 4(a)–(h) reveal the presence of an apatitic phase (loosely matching sium content to 1 wt.% (for heat-treated sample 1Mg–CaP) resulted
ICDD pattern 09-432 for HA). The broadness and low intensity of the in appearance of a very small amount of a new phase as seen in the
peaks suggest small particle size and poor crystallinity. It was observed PXRD pattern shown in Fig. 5(d) which was identified to be Mg-βTCP
that with increasing magnesium content PXRD patterns became even (i.e. magnesium stabilized β-TCP phase) [compared to ICDD patterns
more poorly resolved in [Fig. 4(g)–(l)]. The center of the broad peak 70-068 and 13-0404] [22,61–63]. Further increase in nominal mag-
shifted from 31.9° to 30.11° suggesting formation of non-apatitic calci- nesium contents from 1.5 to 2 wt.% for samples 1.5Mg–CaP and
um phosphate phases. 2Mg–CaP resulted in increase in intensity of the peak around 31.4° corre-
Fig. 5 shows the PXRD patterns for samples heat treated at 900 °C sponding to main peak for Mg-βTCP as seen in Fig. 5(e) and (f). Magne-
for 1 h. PXRD patterns corresponding to heat treated samples Ca–P, sium contents of 4 wt.% and 5 wt.% (corresponding to heat treated
0.1Mg–CaP and 0.5Mg–CaP in Fig. 5(a)–(c) respectively, gave a samples 4Mg–CaP and 5Mg–CaP respectively) resulted in formation of
good match to ICDD pattern 09-432 corresponding to phase-pure phase pure Mg–TCP as seen in PXRD patterns shown in Fig. 5(g) and
crystalline HA. Inability to provide long maturation times (N 18 h) (h). For the aforementioned theoretical compositional range of magne-
often leads to non-stoichiometric apatites which decompose at or sium substitution (i.e. from 0.1 wt.% to 5 wt.%) it is suggested that there
above 750 °C into water, HA and TCP phases [60]. In this case however, are two routes of formation of the Mg-βTCP phase. PXRD patterns for
phase-pure hydroxyapatite was observed for magnesium-less (sample as-precipitated Mg–CaP samples 0.1Mg–CaP, 0.5Mg–CaP, 1Mg–CaP,
Ca–P) and low Mg content samples. This therefore elucidates that the 1.5Mg–CaP and 2Mg–CaP (0.1 wt.% to 2 wt.%) in Fig. 4(b)–(f) revealed
microwave irradiation of suspensions greatly accelerates the matura- an amorphous apatitic phase. It is suggested that the Mg-βTCP phase in
tion process. Peaks observed were better resolved as compared to heat-treated forms of these samples [detected in PXRD patterns
PXRD patterns in Fig. 4(a)–(c) (for the same samples in their as- Fig. 5(b)–(f)] evolves from decreased thermal stability (hence partial
precipitated forms) due to improved crystallinity as a result of heat- decomposition into Mg-βTCP) of the magnesium containing apatitic
lattice (first route). The destabilizing effect of Mg is possibly due to lattice
strain introduced as a result of smaller ionic radius of Mg2+ (0.65 Å)
in comparison with Ca2+ (0.99 Å) which favors the hydroxyapatite–
Mg-βTCP transition upon heating [29,50]. Non-apatitic phases were de-
tected for as-precipitated sample 4Mg–CaP and other samples with
higher magnesium content [Fig. 4(g)–(l)]. It is suggested that this amor-
phous non-apatitic phase/s may contain Mg-βTCP in the as-precipitated
form which later crystallizes upon heat-treatment at 900 °C for 1 h
hence showing up as phase-pure crystalline Mg-βTCP in Fig. 5(g) and
(h) for heat treated samples 4Mg–CaP and 5Mg–CaP respectively (second
route) [64,65]. Further increase in nominal magnesium contents to
10 wt.% for sample 10Mg–CaP resulted in appearance of Mg-βTCP and
an additional phase identified as Stanfieldite corresponding to ICDD pat-
tern 23642 [Fig. 5(i)]. Indeed for higher level of magnesium substitutions
in the TCP phase wherein Mg2+ / (Mg2+ + Ca2+) molar ratio is greater
than 14% this phase has been reported (based on amounts added to solu-
tions this ratio is 40% in our case) [66]. Further increase in magnesium
content to 16 wt.% (heat treated sample 16Mg–CaP) resulted in forma-
tion of phase pure Stanfieldite as seen in Fig. 5(j). Sample 22Mg–CaP
Fig. 4. Powder X-ray diffraction patterns of as-precipitated samples (a) Ca–P (b) 0.1Mg–CaP
(nominally 22 wt.% magnesium) was identified as a biphasic mixture of
(c) 0.5Mg–CaP (d) 1.0Mg–CaP (e) 1.5Mg–CaP (f) 2.0Mg–CaP (g) 4Mg–CaP (h) 5Mg–CaP, Stanfieldite and Farringtonite [ICDD # 33-0876, (Mg3(PO4)2] as seen in
(i) 10Mg–CaP (g) 16Mg–CaP (k) 22Mg–CaP and (l) Mg–P. Fig. 5(k). Indeed a reaction carried out in the absence of calcium resulted
N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293 291

in formation of phase pure Farringtonite for sample Mg–P as seen in increase in nominal magnesium content from 5 wt.% to 22 wt.% (for
Fig. 5(l) [67]. samples 5Mg–CaP, 10Mg–CaP, 16Mg–CaP and 22Mg–CaP) resulted in
no discernible change in the FTIR spectra as seen in Fig. 6(h)–(k). The in-
3.4. Fourier transform infrared spectroscopy creased water retention as observed in Fig. 6(h) for sample 5Mg–CaP may
possibly be due to very amorphous nature of the as-precipitated phase.
FTIR spectra for as-precipitated Mg–CaP samples are shown in Indeed, magnesium has been shown to reduce crystallinity of calcium
Fig. 6(a)–(l). For samples Ca–P, 0.1Mg–CaP and 0.5Mg–CaP typical phosphate phases [5]. The FTIR spectrum for sample MgP is shown in
apatitic FTIR spectra were observed as seen in Fig. 6(a–c). The broad Fig. 6(l). Most of its features are very similar to Fig. 6(k) for sample
bands centered at ca. 3450 cm−1 (in the range of 3650–3000 cm− 1) 22Mg–CaP [due to the presence of the Farringtonite phase as shown in
and 1650 cm−1 (1687–1600 cm−1) corresponded to the presence of PXRD patter in Fig. 4(k)]. The doublet peak in the range of 650 cm−1–
absorbed water in HA. The band in the range of 1510–1390 cm− 1 550 cm−1 reappeared corresponding to bending vibration of the P\\O
corresponded to the asymmetric (υ3) stretching of the C\\O bond in bond in O\\P\\O linkages of phosphate group in the Farringtonite
carbonate. This is unsurprising as water used for preparation of phase (sample MgP).
solutions was not degassed prior to reaction. The band observed in Fig. 7 shows FTIR spectra of samples heat treated at 900 °C for 1 h. In
the range of 1110–1040 cm− 1 corresponded to the asymmetric (υ3) Fig. 7(a–b) for samples Ca–P and 0.1Mg–CaP the sharp peak at ca.
stretching of the P\\O bond of phosphate group in HA. The peak at 3567 cm−1 was assigned to the stretching mode (υs) of the hydroxyl
960 cm−1 was attributed to the symmetric stretching (υ1) of the P\\O group of HA that indicated the crystalline nature of HA (heat-treatment
bond of phosphate group in HA. Peak at 874 cm−1 was due to the bending resulted in increase in crystallinity). The band in the range of 2360–
mode (υ2) of the O\\C\\O linkage in carbonate groups present in HA. 1990 cm−1 was attributed to the overtones and combination of v3 and
The position of the C\\O and O\\C\\O bands and peaks (υ2 and υ3) v1 phosphate modes [68]. The band observed in the range of
indicates that B-type carbonate substitution into HA took place. 1510–1380 cm−1 was attributed to the asymmetric stretching (υ3) of
The peak doublet present in the range of 600–560 cm−1 was attributed the C\\O bond of carbonate group in HA. The peaks at 1091 and
to the bending mode (v4) of the O\\P\\O linkage in phosphate group 1034 cm−1 were ascribed to the asymmetric stretching (υ3) of P\\O
of HA. No change was observed in Fig. 6(c)–(f) with increasing nom- bond of phosphate group in HA. The peak at 960 cm−1 was assigned
inal magnesium content from 0.5 to 2 wt.% (for samples 0.5Mg–CaP, to the symmetric stretching (υ3) of the P\\O bond in phosphate
1Mg–CaP 1.5Mg–CaP and 2Mg–CaP). This suggested that for theoret- group. Peak at 873 cm−1 was attributed to the bending mode (υ2) of
ical magnesium substitution levels of up to 2 wt.% the phases were the O\\C\\O linkage in carbonate groups. The presence of carbonate
apatitic (supported by PXRD patterns in Fig. 4(c)–(f)]. Further in- groups in these samples is unsurprising because water was not
crease in nominal magnesium content up to 4 wt.% (4Mg–CaP) degassed prior to use and all experimental steps were carried out in
caused a broadening of the band initially observed in the range of air. These bands have a lower intensity probably due to CO2 loss upon
1110–1040 cm− 1 (corresponding to v3 asymmetric P\\O stretching) heating as compared to those observed in Fig. 6 for as-precipitated sam-
and in disappearance of the peak at 961 cm− 1 corresponding to v1 ples. The peak at 632 cm−1 corresponded to the librational mode (υL) of
stretching of the P\\O bond [see Fig. 4(g) for sample 4Mg–CaP]. the hydroxyl group. This peak was not observed in FTIR spectra for HA in
The peaks of v4 bending mode of the O\\P\\O linkage were observed as a Fig. 6 (for as-precipitated samples). Indeed, this peak has been used as
doublet (at 609 and 565 cm−1) for sample 2Mg–CaP in Fig. 6(f) however an indicator of crystallinity [69]. Peaks at 606 and 582 cm− 1
this doublet was not observed in Fig. 6(g) corresponding to sample corresponded to bending mode (υ4) of the O\\P\\O linkage in phos-
4Mg–CaP (a band was observed instead). The aforementioned broaden- phate groups of HA. The peak at 470 cm−1 was ascribed to the O\\P\\O
ing, disappearance of peaks (non-resolution) and changing of a doublet bending mode (υ2). The sharpness of peaks at 632, 606 and 582 cm−1
peak into a band can possibly be due to similar vibrations of the P\\O indicated a well crystallized HA. No change was observed in
bond and/or O\\P\\O linkages, but from phosphate groups from a dif- Fig. 7(b)–(d) with increasing magnesium content from 0.1 to 1 wt.%
ferent (non-apatitic) calcium phosphate phase. Indeed a non-apatitic (for samples 0.1Mg–CaP, 0.5Mg–CaP and 1Mg–CaP). This indicates
phase was observed in Fig. 4(g) which crystallized into phase pure that magnesium substitution (theoretically calculated) up to 1 wt.%
Mg-βTCP after heat treatment at 900 °C as seen in Fig. 5(g). Further

Fig. 6. Fourier transform infrared spectra in the range of 4000–400 cm−1 for as precipitated Fig. 7. Fourier transform infrared spectra in the range of 4000–400 cm−1 for heat treated
samples (a) Ca–P (b) 0.1Mg–CaP (c) 0.5Mg–CaP (d) 1.0Mg–CaP (e) 1.5Mg–CaP (at 900 °C for 1 h) samples (a) Ca–P (b) 0.1Mg–CaP (c) 0.5Mg–CaP (d) 1.0Mg–CaP
(f) 2.0Mg–CaP (g) 4Mg–CaP (h) 5Mg–CaP (i) 10Mg–CaP (j) 16Mg–CaP (k) 22Mg–CaP (e) 1.5Mg–CaP (f) 2.0Mg–CaP (g) 4Mg–CaP (h) 5Mg–CaP (i) 10Mg–CaP (j) 16Mg–CaP
and (l) Mg–P. (k) 22Mg–CaP and (l) Mg–P.
292 N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293

(measured to be 0.97 wt.%, see Table 2) did not induce any significant [5] A. Bigi, G. Falini, E. Foresti, M. Gazzano, A. Ripamonti, N. Roveri, J. Inorg. Biochem. 49
(1993) 69.
change in the apatite structure. Indeed only a small amount of
[6] A. Bigi, G. Cojazzi, S. Panzavolta, A. Ripamonti, N. Roveri, M. Romanello, K.N. Suarez,
Mg-βTCP phase was observed in the PXRD pattern shown in L. Moro, J. Inorg. Biochem. 68 (1997) 45.
Fig. 5(d) for sample 1Mg–CaP (major phase being HA). [7] N. Rangavittal, A.R. Landa-Canovas, J.M. Gonzalez-Calbet, M. Vallet-Regi, J. Biomed.
A further increase in nominal magnesium content to 1.5 wt.% (for Mater. Res. 51 (2000) 660.
[8] E. Boanini, M. Gazzano, A. Bigi, Acta Biomater. 6 (2010) 1882.
sample 1.5Mg–CaP) resulted in the appearance of new peaks at 989 [9] R.N. Correia, M.C.F. Magalhaes, P.A.A.P. Marques, A.M.R. Senos, J. Mater. Sci. Mater.
and 945 cm−1 attributed to the phosphate groups in Mg-βTCP phase Med. 7 (1996) 501.
[also confirmed in Fig. 5(e)]. The reduction in intensities of the O\\H [10] R. Zapanta LeGeros, Prog. Cryst. Growth Charact. Mater. 4 (1981) 1.
[11] G. Daculsi, R.Z. LeGeros, D. Mitre, Calcif. Tissue Int. 45 (1989) 95.
peak at 3569 cm−1 and υ3 C\\O (1510–1380 cm−1) band and widening [12] F.C.M. Driessens, Ber. Bunsenges. Phys. Chem. 90 (1986) 760.
of the band associated with asymmetric υ3 P\\O stretching (and [13] S. Kannan, A.F. Lemos, J.M.F. Ferreira, Chem. Mater. 18 (2006) 2181.
appearance of new peaks in this range) from Fig. 7(b)–(f) were due to [14] L.A. Bretado, D.A. Cortes, W. Ortega, D. Renteria, Adv. Appl. Ceram. 108 (2009) 194.
[15] I.R. Gibson, W. Bonfield, J. Mater. Sci. Mater. Med. 13 (2002) 685.
successive reduction in amount of HA (and increase in amount of [16] R.K. Rude, H.E. Gruber, H.J. Norton, L.Y. Wei, A. Frausto, J. Kilburn, Bone 37 (2005)
Mg-βTCP) till phase-pure Mg-βTCP was obtained. It was observed 211.
that O\\H peaks at 3569 and 634 cm− 1 were not present in [17] C.M. Serre, M. Papillard, P. Chavassieux, J.C. Voegel, G. Boivin, J. Biomed. Mater. Res.
42 (1998) 626.
Fig. 7(g) for sample 4Mg–CaP (as it is phase-pure Mg-βTCP). In the [18] D.A. Straub, Nutr. Clin. Pract. 22 (2007) 286.
FTIR spectra shown in Fig. 7(g)–(h) for samples 4Mg–CaP and 5Mg– [19] Y.H. Yun, Z.Y. Dong, Z.Q. Tan, M.J. Schulz, Anal. Bioanal. Chem. 396 (2010) 3009.
CaP (identified earlier as phase-pure Mg-βTCP) the peaks correspond- [20] H. Aydin, O. Deyneli, D. Yavuz, H. Gozu, N. Mutlu, I. Kaygusuz, S. Akalin, Biol. Trace
ing to O\\P\\O υ2 bending at 600 and 552 cm− 1 were observed as a Elem. Res. 133 (2010) 136.
[21] G. Qi, S. Zhang, K.A. Khor, S.W. Lye, X. Zeng, W. Weng, C. Liu, S.S. Venkatraman, L.L.
doublet. This doublet was not observed in Fig. 7(i)–(j) [for samples Ma, Appl. Surf. Sci. 255 (2008) 304.
10Mg–CaP and 16Mg–CaP] possibly due to the bending vibrations of [22] S. Kannan, I.A.F. Lemos, J.H.G. Rocha, J.M.F. Ferreira, J. Solid State Chem. 178 (2005)
the O\\P\\O linkages in phosphate groups in Stanfieldite phase. The 3190.
[23] B. Tomazic, M. Tomson, G.H. Nancollas, Arch. Oral Biol. 20 (1975) 803.
reappearance of the doublet in Fig. 7(k)–(l) for samples 22Mg–CaP [24] E. Odabasi, M. Turan, A. Aydin, C. Akay, M. Kutlu, Ann. Acad. Med. Singap. 37 (2008)
and MgP can be attributed to the formation of Farringtonite phase. 564.
These FTIR results support the PXRD data presented in Fig. 5 suggesting [25] R.K. Rude, H.E. Gruber, H.J. Norton, L.Y. Wei, A. Frausto, J. Kilburn, Osteoporos. Int. 17
(2006) 1022.
that increased magnesium incorporation in HA leads to its phase de- [26] N. Kanzaki, K. Onuma, G. Treboux, S. Tsutsumi, A. Ito, J. Phys. Chem. B 104 (2000)
composition and formation of non-apatitic phases at higher magnesium 4189.
contents. [27] A. Yasukawa, T. Yokoyama, K. Kandori, T. Ishikawa, Colloids Surf. A Physicochem.
Eng. Asp. 238 (2004) 133.
[28] R. Enderle, F. Gotz-Neunhoeffer, M. Gobbels, F.A. Muller, P. Greil, Biomaterials 26
4. Conclusions (2005) 3379.
[29] A.A. Chaudhry, J. Goodall, M. Vickers, J.K. Cockcroft, I. Rehman, J.C. Knowles, J.A. Darr,
J. Mater. Chem. 18 (2008) 5900.
This work demonstrates that microwave irradiation of pre- [30] G. Grandi, C. Heitz, L.A.D. Santos, M.L. Silva, M. Sant'Ana Filho, R.M. Pagnocelli, D.N.
precipitated suspensions results in magnesium substituted phase-pure Silva, Mater. Res. 14 (2011) 11.
HA (up to 0.5 wt.% Mg content) which is thermally stable up to [31] M. Tomozawa, S. Hiromoto, Y. Harada, Surf. Coat. Technol. 6 (7) (2015) 2787–2796.
[32] F. Ren, Y. Leng, R. Xin, X. Ge, Acta Biomater. 6 (7) (2015) 2365–2862.
900 °C. In using this route, long synthesis durations required were [33] E. Landi, G. Logroscino, L. Proietti, A. Tampieri, M. Sandri, S. Sprio, J. Mater. Sci. Mater.
avoided (N18 h maturation time needed to obtain phase pure thermally Med. 19 (2008) 239.
stable HA under standard co-precipitation route); only 2 h was required [34] S.V. Chiranjeevirao, J. Hemmerle, J.C. Voegel, R.M. Frank, Inorg. Chim. Acta 67 (1982)
183.
to obtain a precipitate that gave phase-pure Mg–HA. It was observed [35] I. Cacciotti, A. Bianco, M. Lombardi, L. Montanaro, J. Eur. Ceram. Soc. 29 (2009) 2969.
that a magnesium content of above 0.5 wt.% (i.e. at 1 wt.% and higher) [36] E. Landi, A. Tampieri, M. Mattioli-Belmonte, G. Celotti, M. Sandri, A. Gigante, P. Fava,
yielded non-apatitic magnesium containing calcium phosphates as G. Biagini, J. Eur. Ceram. Soc. 26 (2006) 2593.
[37] Y. Pan, J.L. Huang, C.Y. Shao, J. Mater. Sci. 38 (2003) 1049.
well. This study constitutes the first study of its kind which focuses on
[38] L.J. Peng, C.Q. Ning, D.Y. Ding, X.Y. Liu, J. Biomater. Tissue Eng. 5 (162) (2015).
such a large spectrum of substituting compositions and fully evaluates [39] S. Haque, I. Rehman, J.A. Darr, Langmuir 23 (2007) 6671.
microwave assisted synthesis (especially from a structural standpoint) [40] M.J. Phillips, J.A. Darr, Z.B. Luklinska, I. Rehman, J. Mater. Sci. Mater. Med. 14 (2003)
as a robust tool for tailorable synthesis of bioceramics with enhanced 875.
[41] A.R. Toibah, I. Sopyan, M. Hamdi, S. Ramesh, 4th Kuala Lumpur International Confer-
potential bioactivity. Endeavors which focus on more ionic substitutions ence on Biomedical Engineering 2008 IFMBE Proceedings 21 (2008) 314–317.
(such as zinc and iron) for a range of biological applications will be [42] H.G. Zhang, Q.S. Zhu, Z.H. Xie, Mater. Res. Bull. 40 (2005) 1326.
reported in due course of time. [43] W.L. Suchanek, K. Byrappa, P. Shuk, R.E. Riman, V.F. Janas, K.S. TenHuisen, J. Solid
State Chem. 177 (2004) 793.
[44] A. Siddharthan, T.S.S. Kumar, S.K. Seshadri, Biomed. Mater. (2009) 4.
Acknowledgments [45] N. Rameshbabu, T.S.S. Kumar, K.P. Rao, Mater. Res. Innov. 14 (2010) 45.
[46] A. Lak, M. Mazloumi, M.S. Mohajerani, S. Zanganeh, M.R. Shayegh, A. Kajbafvala, H.
Arami, S.K. Sadrnezhaad, J. Am. Ceram. Soc. 91 (2008) 3580.
Ministry of Science and Technology, Government of Pakistan is ac- [47] S.J. Kalita, S. Verma, Mater. Sci. Eng. C 30 (2010) 295.
knowledged for its developmental funding (IRCBM PC-1). The Higher [48] M. Ipekoglu, S. Altintas, Biyomut: 2009 14Th National Biomedical Engineering Meeting
Education Commission Pakistan is acknowledged for research (NRPU 2009, p. 439.
[49] H. Arami, M. Mohajerani, M. Mazloumi, R. Khalifehzadeh, A. Lak, S.K. Sadrnezhaad,
AAC - 1834) and developmental (DIASF — CIIT Lahore - 0912252K047) J. Alloys Compd. 469 (2009) 391.
grants. School of Chemical and Materials Engineering (SCME), National [50] I.V. Fadeev, L.I. Shvorneva, S.M. Barinov, V.P. Orlovskii, Inorg. Mater. 39 (2003) 947.
University of Science and Technology (NUST) is acknowledged for its [51] A. Marti, Inj. Int. J. Care Inj. 31 (2000) S33–S36.
[52] A.A. Chaudhry, H. Yan, G. Viola, M.J. Reece, J.C. Knowles, K. Gong, I. Rehman, J.A. Darr,
help regarding SEM imaging. CIIT Lahore is thanked for visiting professor- J. Biomater. Appl. 27 (2012) 79.
ships (IR and JAD). Dr. Umair Manzoor is thanked for his help with EDX [53] A.A. Chaudhry, J.C. Knowles, I. Rehman, J.A. Darr, J. Biomater. Appl. 28 (2013) 448.
analysis. Dr. Zenab Sarfaraz and Ayesha Idrees are thanked for their help [54] A.A. Chaudhry, H. Yan, K. Gong, F. Inam, G. Viola, M.J. Reece, J.B.M. Goodall, I. ur
Rehman, F.K. McNeil-Watson, J.C.W. Corbett, J.C. Knowles, J.A. Darr, Acta Biomater.
with BET surface area analysis. 7 (2011) 791.
[55] R. Nazir, N. Iqbal, A. Samad Khan, A. Akram, A. Asif, A. Anwar Chaudhry, I. Rehman, R.
References Hussain, Ceram. Int. 38 (2012) 457.
[56] N. Iqbal, A. Asif, A.A. Chaudhry, R. Nazir, R. Amin, M. Akram, G.Y. Fan, A. Akram, S.H.
[1] R.H. Doremus, J. Mater. Sci. 27 (1992) 285. Park, R. Hussain, Curr. Appl. Phys. 12 (3) (2011) 755–759.
[2] L.L. Hench, J. Am. Ceram. Soc. 74 (1991) 1487. [57] N.M.D.F. Costa, B.R. Melo, R.T. Brito, G.V.D.O. Fernandes, V.G. Bernardo, E.C. Fonseca,
[3] M.J. Olszta, X. Cheng, S.S. Jee, R. Kumar, Y.Y. Kim, M.J. Kaufman, E.P. Douglas, L.B. M.B. Conz, G.R.A. Soares, J.M. Granjeiro, Mater. Res. 12 (2009) 245.
Gower, Mater. Sci. Eng. R. Rep. 58 (2007) 77. [58] A. Hafiz, K.A. Khalid, A. Yusof, M.A. Azril, A. Shukrimi, M.Y. Nazri, C.A. Aminudin, Z.
[4] E.B. Nery, K.L. Lynch, W.M. Hirthe, K.H. Mueller, J. Periodontol. 46 (1975) 328. Zamzuri, F. Fazan, Eur. Cell. Mater.s 16 (2008) 55.
N.I. Khan et al. / Materials Science and Engineering C 56 (2015) 286–293 293

[59] F.D. Burstein, S.R. Cohen, R. Hudgins, W. Boydston, Plast. Reconstr. Surg. 100 (1996) [66] D. Clement, J.M. Tristan, M. Hamad, P. Roux, J.C. Heughebaert, J. Solid State Chem. 78
869. (1989) 271.
[60] K. Ishikawa, P. Ducheyne, S. Radin, J. Mater. Sci. Mater. Med. 4 (1993) 165. [67] U. Klammert, E. Vorndran, T. Reuther, F.A. Muller, K. Zorn, U. Gbureck, J. Mater. Sci.
[61] S. Kannan, J.M. Ventura, J.M.F. Ferreira, Ceram. Int. 33 (4) (2015) 637–641. Mater. Med. 21 (2010) 2947.
[62] S. Kannan, A.F. Lemos, J.H.G. Rocha, J.M.F. Ferreira, J. Am. Ceram. Soc. 89 (2006) 2757. [68] S. Koutsopoulos, J. Biomed. Mater. Res. 62 (2002) 600.
[63] S. Kannan, J.H.G. Rocha, J.M.F. Ferreira, J. Mater. Chem. 16 (2006) 286. [69] A.A. Chaudhry, S. Haque, S. Kellici, P. Boldrin, I. Rehman, A.K. Fazal, J.A. Darr, Chem.
[64] S.L. Rowles, Bull. Soc. Chim. Fr. (1968) 1797. Commun. (2006) 2286.
[65] R.Z. LeGeros, A.M. Gatti, R. Kijkowska, D.Q. Mijares, J.P. LeGeros, Mg-Substituted
Tricalcium Phosphates: Formation and Properties, 2004.

You might also like