Odonnell2008 - Wet - Chemical

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Available online at www.sciencedirect.

com

Acta Biomaterialia 4 (2008) 1455–1464


www.elsevier.com/locate/actabiomat

Structural analysis of a series of strontium-substituted apatites


M.D. O’Donnell *, Y. Fredholm, A. de Rouffignac, R.G. Hill
Imperial College London, Department of Materials, South Kensington Campus, Exhibition Road, London SW7 2AZ, UK

Received 6 November 2007; received in revised form 26 March 2008; accepted 16 April 2008
Available online 3 May 2008

Abstract

A series of Sr-substituted hydroxyapatites, (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00, were made by a standard
wet chemical route and investigated using X-ray diffraction (XRD), Rietveld refinement and Raman spectroscopy. We report apatites
manufactured by two synthesis routes under 90 °C, and only the fully Sr-substituted sample had a small amount of an impurity phase,
which is believed to be strontium pyrophosphate. Lattice parameters (a and c), unit cell volume and density were shown to increase lin-
early with strontium addition and were consistent with the addition of a slightly larger and heavier ion (Sr) in place of Ca. XRD Lorentz-
ian peak widths increased to a maximum at x = 0.50, then decreased with increasing Sr content. This indicated an increase in crystallite
size when moving away from the x = 0.50 composition (d  9.4 nm). There was a slight preference for strontium to enter the Ca(II) site
1
in the mixed apatites (6 to 12% depending on composition). The position of the Raman band attributed to v1 PO3 4 at around 963 cm in
1
hydroxyapatite decreased linearly to 949 cm at full Sr-substitution. The full width at half maximum of this peak also correlated well
and increased linearly with increasing crystallite size calculated from XRD.
Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Raman; Apatite; Strontium; XRD; Rietveld

1. Introduction samples are typically only analysed by X-ray diffraction


(XRD). Heijligers et al. [1] performed a structural study
It is known that strontium and other divalent cations of Sr–Ca apatites made by solid state synthesis using
with a similar charge-to-size ratio to calcium can readily XRD but had significant tristrontium phosphate present
substitute in the lattice of hydroxyapatite. Also it has been for compositions above 80% Sr-substitution. Bigi et al.
reported that there is immiscibility in this system and com- looked at the structure of 0, 20 and 60% substitution of
plete solid-solution cannot occur, although this may be due Sr for Ca in hydroxyapatite [8].
to poor sample preparation [1]. Apatites have many appli- Raman microspectrometry is powerful investigative
cations, from optoelectronics [2] to waste immobilization tool, as the probe is 10–100 times smaller than similar
[3] and biomaterials [4]. Strontium is also beneficial for bio- micro-infrared spectroscopic techniques. It is useful for
logical applications for bone regeneration due to the recent biological samples due to its non-invasive and non-destruc-
success of treatments such as strontium ranelate [5], which tive nature, and fluorescence is minimized. No special sam-
stimulates bone formation, decreases bone resorption and ple preparation is required as the technique is performed in
reduces the risk of vertebral fractures in postmenopausal reflection and hence artefacts from the preparation process
osteoporosis. Both calcium- and strontium-hydroxyapatite are eliminated. Complex studies can be performed, such as
are hexagonal (space group P63/m) [6,7]. No complete sys- at interfaces and line-scans of cross-sections. The func-
tematic studies have been performed on this system manu- tional groups present in biological and synthetic apatites,

factured by a wet chemical route to our knowledge, and such as phosphate (PO3 4 ), hydroxyl (OH ), acidic phos-
phate (HPO4 ), carbonate (CO3 ) and fluoride (F), can
2 2

*
Corresponding author. Tel.: +44 20 7594 6814; fax: +44 20 7594 6757. be probed, and peak width, position and multiplicity corre-
E-mail address: m.odonnell@imperial.ac.uk (M.D. O’Donnell). lated with composition and structure.

1742-7061/$ - see front matter Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actbio.2008.04.018
1456 M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464

The purpose of this study was to investigate the crystal- Solution A–solution B mix. Solution B was added drop-
line structure of a series of apatites, where Ca is progres- wise to the constantly stirred solution A over a period of
sively substituted for Sr, by XRD, Rietveld refinement 2 h. The pH of the mixture was adjusted to about 11 with
and Raman spectroscopy. a 30% ammonia solution, if needed. The precipitate formed
was stirred for 1 h, before being left overnight. The precip-
2. Experimental itate was centrifuged and washed twice with distilled water.
Finally, the precipitate was dried for 20 h at 85 °C. The
2.1. Synthesis dried material was powdered using a Gyro-mill and sieved
with a 38 lm sieve.
Strontium-substituted hydroxyapatites were synthesized
by a wet chemical route. Two reaction routes were used [9]. 2.2. XRD

2.1.1. Route 1 A Phillips powder diffractometer with a copper (Cu Ka)


Solution A preparation. Calcium nitrate 4-hydrate X-ray source (Philips PW 1700 series diffractometer, Phi-
(Ca(NO3)24H2O) and/or strontium nitrate (Sr(NO3)2) were lips, Eindhoven, the Netherlands) was used to characterize
dissolved in 200 ml of distilled water. Table 1 gives the the glass samples. The powdered samples (<38 lm particle
weights used for each apatite composition. The pH of the size) were scanned between 2h values of 10 and 80° with a
solution was adjusted to 11 using 30% ammonia solution, step size of 2h = 0.04° in order to determine the crystal
then 400 ml of distilled water was added to the solution. structure of each apatite. This step size results in a d-spac-
The solution was constantly stirred. ing deviation of around 0.009 Å at 20° 2h, consistent with
Solution B preparation. Diammonium hydrogen ortho- the errors reported in the refinement below. These param-
phosphate ((NH4)2HPO4) was dissolved in 120 ml of dis- eters gave sufficient resolution for the refinement, with the
tilled water. The amount of (NH4)2HPO4 used is given most intense peaks occurring in the 2h = 30–35° region.
in Table 1. The pH of the solution was adjusted to 11 XRD was then carried out and the results analysed using
using ammonia solution; however, no hydroxides precip- software containing a database of standard diffraction files.
itated, which have lower solubility than amide salts, tri- Rietveld refinement was performed with GSAS [10] and
calcium phosphate and apatite. Distilled water (160 ml) EXPGUI [11] software using a 10-term shifter-Chebyschev
was then added to the constantly stirred solution. The background function. Initial atomic coordinates and unit
solution was filtered using a Buchner funnel, as cell dimensions were taken from previously published data
described [9]. [6,7]. Parameters varied in the refinement were atomic
Solution A–solution B mix. Solution B was added drop- coordinates (x, y and z), unit cell parameters (a and c), dis-
wise to the constantly stirred solution A over a period of placement parameters (Uiso), peak intensity scaling and
approximately 2 h. The pH was kept at around 11. A pre- peak profile parameters related to particle size and stress.
cipitate formed and was stirred for 1 h, before being left For the mixed apatites, the fractional occupancies of both
overnight. The precipitate was centrifuged and washed divalent cation sites were initially set at the stoichiometric
twice with distilled water. Finally, the precipitate was dried values, then refined to total unity as there is some evidence
for 20 h at 85 °C. The dried material was crushed using a that strontium preferentially substitutes onto the Ca(II) site
Gyro-mill and sieved with a 38 lm sieve into coarse and [8]. To determine the Bragg peak widths, the Bragg reflec-
fine fractions. tion at around 2h = 25.9°, corresponding to the 0 0 2 hkl
reflection, was Lorentzian deconvoluted with Microcal Ori-
2.1.2. Route 2 gin software.
Solution A preparation. Calcium hydroxide (Ca(OH)2)
and strontium hydroxide (Sr(OH)2) were dissolved in 2.3. Raman
500 ml of distilled water. Table 2 gives the weights used
for each apatite composition. The Raman system used consisted of a Renishaw RM
Solution B preparation. Orthophosphoric acid solution 2000 spectrometer (Renishaw LC, UK) connected to a
(85% H3PO4) was diluted in 500 ml of distilled water. Leica microscope with an objective (50). A 785 nm
The amount of H3PO4 used is given in Table 2. 300 mW line focus laser with 100 mW power at the sample

Table 1
Experimental weights (in g) for route 1 synthesis for the series (SrxCa1x)5(PO4)3OH, where x = 0.00 and 1.00
Sr (x) Ca(NO3)2  4H2O wt.% mol.% Sr(NO3)2 wt.% mol.% (NH4)2HPO4 wt.% mol.%
1.00 – – – 42.34 72.76 62.50 15.85 27.24 37.50
0.50 23.62 38.95 31.25 21.17 34.91 31.25 15.85 26.14 37.50
0.25 35.42 57.27 46.87 10.58 17.11 15.62 15.85 25.63 37.51
0.00 47.23 74.87 62.50 – – – 15.85 25.13 37.50
M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464 1457

Table 2
Experimental weights (in g) for route 2 synthesis for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00
Sr (x) Ca(OH)2 wt.% mol.% Sr(OH)2 wt.% mol.% H3PO4 wt.% mol.%
1.00 – – – 30.10 64.36 59.26 16.67 35.64 40.74
0.75 4.58 10.45 14.80 22.58 51.52 44.46 16.67 38.03 40.74
0.00 18.34 52.39 59.27 – – – 16.67 47.61 40.73

was used. This wavelength and 100% power was sufficient Overlapping peak positions were obtained by Lorentzian
to produce a high signal to noise ratio Raman spectrum. deconvolution using Microcal Origin software.
The laser spot size was calculated to be 10 lm wide by
20 lm high and 40 lm deep. The spectrometer was set up 3. Results and discussion
with a spectrometer slit of 50 lm and 8 CCD (charge-cou-
pled device) pixels. The powdered samples were placed on a 3.1. XRD and Rietveld refinement
glass slide to collect the spectra. Ten 10 s spectra were
taken between Raman wave number shifts of 100 and Fig. 1a shows the XRD traces of the Sr-substituted apa-
1500 cm1. Quartz was used as a calibration material with tites and Rietveld refinement. Traces were shifted on the
the main Raman resonance peak at around 521 cm1. y-axis for display purposes. The fitting was good for all

a 3500

3000

2500
Intensity (a.u.)

2000

1500

1000

500

0
10 20 30 40 50 60 70 80
o
2-theta /

x=0.00 x=0.25 x=0.50 x=0.75 x=1.00

600

500

400
Intensity

300

200

100

-100
10 20 30 40 50 60 70 80
o
2-theta /
Data Fit Background Residual

Fig. 1. (a) XRD traces and Rietveld refinement of the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00: data points (XRD) and solid
lines (Rietveld); (b) Rietveld refinement for x = 0.50, with residual and positions of Bragg reflections.
1458 M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464

samples, as demonstrated in the high R2 values (all >0.9, on pure Sr-hydroxyapatite and mixed Ca–Sr-hydroxyapa-
not shown) indicating excellent correlation. Fig. 1b shows tite experimental densities, Fig. 5 shows the density of a
a typical fit for the series for the x = 0.5 sample, also show- Yb-doped Sr-fluorapatite [2]. As F and OH have
ing the residual and background. The diffraction peaks approximately the same mass and ionic radii, the density
shift to lower 2h values with Sr-addition, indicating an of these materials should not differ greatly from Sr-
increase in d-spacings and hence lattice parameters. This hydroxyapatite. As Fig. 5 shows, there is good agreement
would be expected as Sr is slightly larger than Ca (118 between the published and Rietveld data.
and 100 pm for octahedral coordination [12]). Peak inten- Crystal sizes, d, were calculated from the full width at
sity also increases with Sr-addition. This is to be expected half maximum values (FWHMs) of the most intense Bragg
as Sr is heavier and contains more electrons than Ca, and reflection at around 2h = 25.9°, corresponding to the 0 0 2
will therefore more effectively scatter X-rays, so there is a hkl reflection and using the formula equation shown below
general increase in crystallinity with Sr-addition (see in Eq. (1) [16]
below). 0:9k
Lattice parameters (a and c) increase linearly with Sr- d¼ ð1Þ
w cos hx
addition, as shown in Figs. 2 and 3, consistent with a larger
ion entering the apatite lattice. As a result of this, the unit where k is the wavelength of the X-rays (1.5406 Å), w is
cell volume also increases linearly, as seen in Fig. 4. Density the FWHM of the Bragg peak and hx the angle of the
calculated from the Rietveld refinement in Fig. 5 also Bragg reflection. The crystallite size decreased to a mini-
increases linearly with Sr-addition due to the replacement mum at around x = 0.5 then increased with Sr content
of a heavier ion for Ca in the crystal structure (87.62 and (Fig. 6). This is consistent with the observations of Li
40.078 g mol1, respectively [12]), and the Ca-rich end of et al. [17], who showed, with a series of three Sr–Ca apa-
the series agree well with previously published experimental tites (0.3, 1.5 and 15% Sr-substitution), that crystallite
data on pure Ca-hydroxyapatites made by aqueous gel size decreases as Sr-substitution approaches 15%. This
casting [13] and sintering [14] and a low Sr-substitution study did not report higher Sr-substitutions. It has been
Ca-hydroxyapatite [15]. Due to a lack of published data indicated that strontium addition to hydroxyapatite

Table 3
Parameters from Rietveld refinement of the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00, with standard deviations
Sr Ca a / Å 2ra / Å c / Å 2rc / Å V / Å3 2rV / Å3 q/g.cm3 d / nm
0.00 1.00 9.411 0.004 6.877 0.003 527.5 0.378 3.163 17.55
0.25 0.75 9.505 0.007 6.950 0.006 543.8 0.648 3.384 11.37
0.50 0.50 9.596 0.004 7.054 0.004 562.6 0.452 3.692 9.38
0.75 0.25 9.659 0.005 7.182 0.004 580.3 0.740 3.897 12.94
1.00 0.00 9.777 0.003 7.288 0.003 603.3 0.472 4.068 20.50

9.80

9.75
y = 0.35x + 9.41
2
9.70 R = 0.99

9.65
a /Å

9.60

9.55

9.50

9.45

9.40
0.0 0.2 0.4 0.6 0.8 1. 0
Fraction Sr

Fig. 2. Variation in lattice parameter, a, for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00; error bars: +2r from Table 3.
M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464 1459

7.30

7.25 y = 0.42x + 6.86


2
R = 0.99
7.20

7.15

7.10
c /Å

7.05

7.00

6.95

6.90

6.85
0.0 0.2 0.4 0.6 0.8 1.0
Fraction Sr

Fig. 3. Variation in lattice parameter, c, for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00; error bars: ±2r from Table 3.

610
y = 75.25x + 525.88
600 2
R = 1.00
590

580

570
V / Å3

560

550

540

530

520
0.0 0.2 0.4 0.6 0.8 1.0
Fraction Sr

Fig. 4. Variation in unit cell volume for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00; error bars: ±2r from Table 3.

4.2

y = 0.93x + 3.18
4.0 2
R = 0.99

3.8
-3
ρ / g.cm

3.6

3.4

3.2

3.0
0.0 0.2 0.4 0.6 0.8 1.0
Fraction Sr
Rietveld Literature

Fig. 5. Variation in density for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00; error bars ±1%.
1460 M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464

22
2
y = 37.73x - 34.74x + 17.57
20 2
R = 1.00
18

16
d / nm

14

12

10

8
0.0 0.2 0.4 0.6 0.8 1.0
x

Fig. 6. Variation in crystallite size from XRD Lorentzian peak widths for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00.

Table 4 1.00
Occupancies of M(I) and M(II) sites from Rietveld refinement
Site x (Sr)
Occupancy - Rietveld

0.00 0.25 0.50 0.75 1.00 0.75


Ca(I) 1.00 0.81 0.52 0.31 0.00
Sr(I) 0.00 0.19 0.48 0.69 1.00
Ca(II) 1.00 0.75 0.45 0.20 0.00 0.50
Sr(II) 0.00 0.25 0.55 0.80 1.00
Mean Sr 0.00 0.22 0.52 0.74 1.00
0.25

broadens the crystal size distribution, which contributes


0.00
to the improved mechanical strength of bone [17]. When
0.00 0.25 0.50 0.75 1.00
x = 0.5, the Sr–Ca apatite will be at its most disordered;
therefore it would be expected that the crystallite size ap- Sr content (x ) - stoichiometry
proaches a minimum.
Ca(I) Sr(I) Ca(II) Sr(II)
The Rietveld refinement gave a good fit with a slight
preference for strontium occupancy on the Ca(II) site, as 1.00
observed by Bigi et al. [8]. There is a 6 to 12% preference y = 1.01x - 0.01
of Sr for the M(II) site, with increasing preference as the 2
R = 1.00
Sr content (x ) - Rietveld

Sr content increases. The occupancies of the sites can be 0.75


seen in Table 4 and is plotted graphically in Fig. 7a, and
the correlation to the stoichiometric compositions agrees
well, as seen in Fig. 7b. 0.50

3.2. Raman spectroscopy


0.25
Figs. 8–10 show the Raman spectra of the apatites syn-
thesized in this study, normalized to the v1 PO3 4 band at
around 955 cm1 and shifted on the y-axis for presentation
0.00
purposes. It can be seen from the Raman spectra in Fig. 8
0.00 0.25 0.50 0.75 1.00
that there is a general shift to a lower wave number of
Raman bands with Sr-addition (see Table 5). The main Sr content (x ) - stoichiometry
Raman band at around 955 cm1 due to the v1 PO3 4 vibra- Fig. 7. (a) Occupancies of M(I) and M(II) sites and (b) mean Sr
tion decreases linearly with Sr-addition (see Figs. 8, 10 and compositions from occupancies.
M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464 1461

11). As the frequency of a Raman band is dependent on and the strength of the forces between the atoms/ions
lattice vibrations, the masses of the atoms/ions present define the position of the vibration. The frequency of the

Wavelength / nm
810 820 830 840 850 860
3
Normalised intensity (a.u.)

x=1.00
2

x=0.75

x=0.50
1

x=0.25

x=0.00

0
400 600 800 1000 1200
-1
Raman shift / cm

Fig. 8. Raman spectra for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00, with 782 nm excitation.

Wavelength / nm
812 814 816 818 820 822 824 826
0.40

0.35

x=1.00
Normalised intensity (a.u.)

0.30

x=0.75
0.25

x=0.50
0.20

0.15
x=0.25

0.10
x=0.00

0.05
400 450 500 550 600 650
-1
Raman shift / cm

Fig. 9. Raman spectra forth series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00, with 782 nm excitation (400–650 cm1 region).
1462 M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464

Wavelength / nm
845 850 855 860
2.5

2.0
Normalised intensity (a.u.)

x=1.00

1.5

x=0.75

x=0.50
1.0

x=0.25
0.5

x=0.00

0.0
900 950 1000 1050 1100 1150
-1
Raman shift / cm

Fig. 10. Raman spectra for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00, with 782 nm excitation (900–1150 cm1 region).

Table 5
Raman band positions (in cm1) for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00
x v2 PO3
4 v4 PO3
4 P–O–P v1 PO3
4 v3 PO3
4 O–P–O
0.00 430, 446 580, 592, 609 – 963 1031, 1048, 1070 –
0.25 426, 447 579, 589, 606 – 958 1022, 1048, 1069 –
0.50 424, 443 582, 601 – 956 1034, 1051, 1064 –
0.75 425, 442 576, 585, 599 – 952 1027, 1044, 1065 –
1.00 422, 442 573, 581, 595 714, 738 949 1023, 1051, 1058 1051, 1058

965
y = -13.16x + 961.98
2
Position of PO4 v1 band / cm-1

R = 0.99
960

955

950

945
0.0 0.2 0.4 0.6 0.8 1.0
x
Fig. 11. Change in the v1 PO3
4 Raman band position for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75 and 1.00.
M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464 1463

18

16 y = -0.94x + 26.34

FWHM of PO4 v1 band / cm-1


2
R = 0.90

14

12

10

6
8 10 12 14 16 18 20 22
d / nm
Fig. 12. Change in the FWHM of v1 PO3
4 Raman band plotted against crystallite size for the series (SrxCa1x)5(PO4)3OH, where x = 0.00, 0.25, 0.50, 0.75
and 1.00.

vibration, f, can be described classically by a harmonically bone) relative to the v1 phosphate peak, and any carbonate
oscillating ball and spring model, shown in Eq. (2), known contamination from the reactants would be small. These
as the Szigeti relationship [18] features could be due to a small amount strontium pyro-
sffiffiffi phosphate impurity (Sr2P2O7), as Raman bands have been
1 k previously reported in a series of phosphate glasses at
f ¼ ð2Þ
c l around 775 and 1070 cm1, identified as P–O–P and O–
P–O vibrations in P2 O4 7 dimmers, respectively [21], with
where c is the velocity of light in a vacuum the latter band higher in intensity, as seen here. This impu-
(2.998108 m s1), k is the bond force constant (typically rity is likely to be minor only, as one peak can be seen that
in N m1) and l is the reduced mass of the two bonding may correspond to this phase at 2h = 33.04° in Fig. 1 for
atoms (equal to m1m2/(m1 + m2)). As the bond strengths x = 1.00. This may be due to the 3 2 0 hkl reflection and
of Ca–O and Sr–O are similar (bond enthalpies of 402.1 is the third most intense peak in Sr2P2O7 (space group
and 425.5 kJ mol1, respectively [19]), k would be expected Pmna) [22]. The highest and second highest intensity peaks
to be similar for both bonds. Therefore, as Sr is substituted are likely to be hidden in the strong apatite peaks.
for Ca, l will increase, shifting f, and hence the Raman Finally, the FWHM of the v1 PO3 band was plotted
4
shift to lower values as PO34 is associated with the heavier against crystallite size from Table 3 and Fig. 6, and is
Sr cation. In practice, this shift could be used to determine shown in Fig. 12. There is a good correlation between
the change in Sr in apatite compositions in future in vivo crystal size and peak width, with linear broadening with
and in vitro studies, in implant materials for example. decreasing size, regardless of the composition. Again,
The following bands can be seen in all samples (Figs. 8– this could be used to estimate disorder in apatite
10) and are tabulated in Table 5 [20]: two bands at around samples.
430 cm1, attributed to v2 PO3 4 ; three bands around
590 cm1, attributed to v4 PO3 4 (except the x = 0.50 sam-
ple, where only two bands could be resolved); one band 4. Conclusions
at around 955 cm1, attributed to v1 PO3 4 ; and three bands
around 1050 cm1, attributed to v3 PO3 4 . In addition, two  A series of Sr-substituted hydroxyapatites were success-
bands were seen at around 725 cm1 for the pure Sr- fully synthesized by two routes under 90 °C with only
hydroxyapatite (x = 1.00). Unusually, for this sample the the fully substituted material containing a small impu-
intensity of the bands in the region of the normally weak rity, likely to be Sr2P2O7.
v3 PO34 vibrations were approximately half the strongest  Lattice parameters, unit cell volume and density
v1 PO34 band. This spectral region contains the carbonate decreased linearly with strontium content, consistent
vibration in apatite: 1100 cm1 for A-type carbonate with addition of a larger, heavier ion.
(OH) and 1070 cm1 for the more common B-type car-  Crystallite size decreased with up to 25% substitution of
bonate (PO3 4 ) [20]. However, these vibrations are weak Sr then increased with higher content.
in biological apatites, which typically show carbonate con-  There was a slight preference for Sr-occupancy of the
tents of between 2 (dental enamel) and 8% (dentine and M(II) site (6 to 12%) in agreement with previous studies.
1464 M.D. O’Donnell et al. / Acta Biomaterialia 4 (2008) 1455–1464

 The v1 PO3 4 Raman vibration at around 963 cm1 [7] Sudarsanan K, Young RA. Structure of strontium hydroxide
decreased linearly with Sr-addition, indicating that this phosphate, Sr5(PO4)3OH. Acta Crystallogr B 1972;28:3668–70.
[8] Bigi A, Falini G, Gazzano M, Roveri N, Tedesco E. Structural
could be used to accurately determine compositions of refinements of strontium substituted hydroxylapatites. Mater Sci
solid-solution compounds. Forum 1998;278(2):814–9.
 The FWHM of the v1 PO3 4 Raman band decreased lin- [9] Best SM, Bonfield W, Gibson IR, Jha LJ, Santos JDDS, inventors,
early with crystallite size, which could also be used to Abonetics Limited, assignee. Silicon-substituted apatites and process
determine disorder in chemically dissimilar apatite for the preparation thereof. United States Patent No. 6312468; 2001.
[10] Larson AC, Von Dreele RB. General structure analysis system
systems. (GSAS). Los Alamos National Laboratory; 1994.
[11] Toby BH. EXPGUI, a graphical user interface for GSAS. J Appl
Crystallogr 2001;34:210–3.
Acknowledgements [12] Greenwood NN, Earnshaw A. Chemistry of the elements. New
York: Pergamon Press; 1984.
MDO would like to thank Dr. Molly Stevens, Dr. Gavin [13] Chen B, Zhang Z, Zhang J, Dong M, Jiang D. Aqueous gel-casting of
Jell and Mr. Robin Swain at IC for use of the Raman hydroxyapatite. Mater Sci Eng A 2006;435–436(5):198–203.
[14] Gross KA, Rodrı́guez-Lorenzo LM. Sintered hydroxyfluorapatites.
equipment and Dr. Stephen Skinner at IC for proofreading Part I. Sintering ability of precipitated solid solution powders.
the manuscript. Biomaterials 2004;25(7–8):1375–84.
[15] Kikuchi M, Yamazaki A, Otsuka R, Akao M, Aoki H. Crystal
structure of Sr-substituted hydroxyapatite synthesized by hydrother-
References mal method. J Solid State Chem 1994;113(2):373–8.
[16] Cullity BD. Elements of X-ray diffraction. Reading, MA: Addison-
[1] Heijligers HJM, Driessens FCM, Verbeeck RMH. Lattice-parameters Wesley; 1959.
and cation distribution of solid-solutions of calcium and strontium [17] Li ZY, Lam WM, Yang C, Xu B, Ni GX, Abbah SA, et al. Chemical
hydroxyapatite. Calcified Tissue Int 1979;29(2):127–31. composition, crystal size and lattice structural changes after incor-
[2] Deloach LD, Payne SA, Smith LK, Kway WL, Krupke WF. Laser poration of strontium into biomimetic apatite. Biomaterials
and spectroscopic properties of Sr5(PO4)3F–Yb. J Opt Soc Am B: Opt 2007;28(7):1452–60.
Phys 1994;11(2):269–76. [18] Parker JM, Seddon AB. Infrared-transmitting optical fibres. In:
[3] Ma QY, Traina SJ, Logan TJ, Ryan JA. In situ lead immobilization Cable M, Parker JM, editors. High-performance glasses. Lon-
by apatite. Environ Sci Technol 1993;27(9):1803–10. don: Blackie; 1992. p. 252–86.
[4] Ohtsuki C, Kokubo T, Yamamuro T. Mechanism of apatite [19] Kerr JA. CRC handbook of chemistry and physics. 81st ed. Florida
formation on CaO–SiO2P2O5 glasses in a simulated body-fluid. J (FL): CRC Press; 2000.
Non-Cryst Solids 1992;143(1):84–92. [20] Penel G, Leroy G, Rey C, Bres E. MicroRaman spectral study of the
[5] Meunier PJ, Roux C, Seeman E, Ortolani S, Badurski JE, Spector PO4 and CO3 vibrational modes in synthetic and biological apatites.
TD, et al. The effects of strontium ranelate on the risk of vertebral Calcified Tissue Int 1998;63(6):475–81.
fracture in women with postmenopausal osteoporosis. N Engl J Med [21] Santos LF, Almeida RM, Tikhomirov VK, Jha A. Raman spectra
2004;350(5):459–68. and structure of fluoro aluminophosphate glasses. J Non-Cryst Solids
[6] Sudarsanan K, Young RA. Significant precision in crystal structural 2001;284(1–3):43–8.
details: holly springs hydroxyapatite. Acta Crystallogr B [22] Barbier J, Echard JP. A new refinement of alpha-Sr2P2O7. Acta
1969;25:1534–43. Crystallogr C 1998;54:2.

You might also like