The XRD results of powder of strontium-magnesium-doped fluorapatite Ca8MgSr(PO4)6F2 obtained by co-precipitation of solutions of the starting components are shown in Fig. 1. An analysis of the diffraction patterns of the obtained powders indicates that, as a result of chemical precipitation, single-phases of fluorapatites Ca8MgSr(PO4)6F2 were produced in a good agreement with previous results for Ca10(PO4)6F2 powders [2]. XRD patterns line demonstrate a significant broadening for Ca8MgSr(PO4)6F2 powder. This fact is an evidence of small sizes of coherent scattering regions (CSR). The substructure characteristics analysis has shown that the size of CRS for the fluorapatites was in the range of 30-33 nm.
These results are in good agreement with the data of the previous study [38]. As can be seen in transparent electron microscopy (TEM) images (Fig.2), powders of Ca10(PO4)6F2 fluorapatite synthesized by chemical precipitation method were agglomerated and mainly consisted from the nano crystalline particles with the average size of 20-30 nm (Fig. 2a). The Ca8MgSr(PO4)6F2 fluorapatite powders produced by chemical precipitation method were also agglomerated with larger particles of the average size of 50-60 nm. (Fig. 2b).
As it is shown [39], ionic radius of the Sr2+ (1.13 Å) is slightly bigger than of the Ca2+ (0.99 Å) one. On the contrary, ionic radius of the Mg2+ (0.65 Å) is significantly smaller than of Ca2+, and ionic substitution of the calcium by magnesium ions leads to noticeable decrease in lattice parameters of fluorapatite. At present study, the substitutions of the Ca2+ by the Sr2+ and Mg2+ ions in the synthesized Ca10(PO4)6F2 powders leaded to the increasing of fluorapatite lattice parameters: a = 9.393Å; c = 6.894Å.
The differential thermal and thermogravimetric analysis (DTA/TGA) curves of powders Ca10(PO4)6F2 and Ca8MgSr(PO4)6F2 produced by chemical co-precipitation have a similar character (Fig. 3). The obtained data indicate the presence of two thermal effects. The first one relates to the removal of adsorbed water at a temperature of ~ 60-110 °C. The second may be associated with the thermal decomposition of fluorapatites by the reaction [40]:
Ca10(PO4)6F2 → 3 Ca3(PO4)2 + CaF2.
For fluorapatite, the partial decomposition temperature was near 900 °C (Fig 3a). The partial decomposition temperature of fluorapatite Ca8MgSr(PO4)6F2, however, decreases to 650 °C, when calcium is replaced by strontium and magnesium (Fig 3b).
The phase composition of sintered (1250 °C, 6 hours) strontium-magnesium-doped fluorapatite sample was represented by the main phase Ca8MgSr(PO4)6F2 (Fig. 4). The weight content of the Ca8MgSr(PO4)6F2 phase was about 81.6 wt%. The presence of tricalcium phosphate Ca3(PO4)2 concomitant phase for strontium-magnesium-doped fluorapatite sample as a result of the process of Ca substitution by Mg and Sr ions was detected.
According to the density measurements (rap), it is found that the maximum relative densities of fluorapatites Ca10(PO4)6F2 and Ca8MgSr(PO4)6F2 were recorded after sintering at a temperature of 1250 °C are 92% and 94% (2.95 and 3.14 g/cm3), respectively. A slight increase in the density of the Ca8MgSr(PO4)6F2 sample compared to Ca10(PO4)6F2 is possibly due to the presence of a some amount of calcium fluoride CaF2 [40] and tricalcium phosphate Ca3(PO4)2. As known from the phase diagram of the calcium fluoride and the tricalcium phosphate, an eutectic between CaF2 and Ca10(PO4)6F2 phases is formed near the temperature of 1200 °C. This effect results in liquid phase formation and more intensification of the sintering process for the Ca8MgSr(PO4)6F2 fluorapatite. The observed increase of density of strontium-magnesium-doped fluorapatite in comparison to fluorapatite samples may be related to liquid phase sintering of the Ca8MgSr(PO4)6F2 fluorapatite.
SEM images of cleaved fluorapatite samples confirm the homogeneous structure of fluorapatite and strontium-magnesium-doped fluorapatite, see Fig. 5. As can be seen, a large number of small pores are presented on the cleaved surface of Ca10(PO4)6F2, while a smaller number of pores, but with larger dimensions, are observed on the cleaved surface of Ca8MgSr(PO4)6F2. This difference is attributed to the realization of the liquid-phase sintering mechanism during heat treatment of the Ca8MgSr(PO4)6F2 sample at the temperature of 1250 °C for 6 hours.
The EDX analysis data of the selected sites of cleaved surface of the Ca10(PO4)6F2 and Ca8MgSr(PO4)6F2 samples are presented in Fig 6. The peaks of Ca, P, O, F, Sr and Mg correspond to the peaks of the main fluorapatites elements. The EDX analysis data have shown some deviations from the calculated values of the consisting elements of both fluorapatite (Ca = 34.15; Sr = 8.29; P = 17.61; O = 36.35; F = 3.60 wt.%) and strontium-magnesium fluorapatite (Ca = 30.82; Sr = 8.42; Mg = 2.33; P = 17.88; O = 36.9; F = 3.65 wt.%).
Fig. 7 shows the data of IR spectroscopy of fluorapatite samples Ca10(PO4)6F2 and Ca8SrMg(PO4)6F2 after sintering. The fluorapatite spectrum contains a series of intense bands with narrow, clear maximum which are typical behavior for the fluorapatite. The spectrum demonstrates the bands of 3450 and 475 cm–1, associated with the presence of adsorbed H2O given in Table 1. In contrast to fluorapatite, a number of differences were observed in the spectrum of strontium-magnesium-doped fluorapatite:
- a characteristic peak of 725 cm-1 appears, corresponding to symmetric vibrations of the bridging bonds of the P-O-P diorthogroups. This indicates the association of phosphate tetrahedra;
- the band disappears in the region of 960 cm-1. This is caused by the degeneracy of the stretching vibrations of the PO43- ion due to a change in the coordination environment and symmetry of the PO43- ion due to the breaking of bonds between the phosphate tetrahedron and calcium ions;
- significant reduction in the intensity of all bands for the spectrum associated with both the structure of fluorapatite and adsorbed water.
There was detected, that the Sr and Mg doping of fluorapatite results in reduction of the band intensity and band resolution corresponding to PO4−3 vibration modes. This fact has suggested that Ca is substituted by Sr and Mg in the fluorapatite lattice.
Table 1. Assignment of the bands in the IR spectra of fluorapatite samples
Ca10(PO4)6F2
|
Сa8SrMg(PO4)6F2
|
Bands [41]
|
475 cm-1
|
460 cm-1, 480 cm-1.
|
Vibrations of OH- groups replacing F- in apatite structure
|
575 cm-1
|
575 cm-1
|
Bending vibrations of PO43- tetrahedra
|
610 cm-1
|
605 cm-1
|
|
725 cm-1
|
Symmetric vibrations of bridging bonds P-O diorthogroups
|
965 cm-1
|
|
Stretching vibrations of PO43- tetrahedra
|
1040 cm-1
|
1040 cm-1
|
1090 cm-1
|
1100 cm-1
|
3450 cm-1 (~10%)
|
3430 cm-1 (~ 5%)
|
Adsorbed Н2О (stretching vibrations Н-О-Н)
|
In a research related to Raman spectroscopic images of bones [42] it has been demonstrated that strontium can heterogeneously distribute in bone mineral, with a higher amount in newly formed bone tissue than in old bone tissue. Additionally, it is shown that the strontium incorporation does not make any significant change in the crystal lattice parameters. In this study, in accordance with the infrared spectroscopy spectra, the Raman spectra of fluorapatite and strontium-magnesium-doped fluorapatites exhibit internal vibrational modes for the group of PO43-, see Fig. 8. Thus, the most intense line of the symmetric stretching vibration ν1(PO) at 961 cm-1, which is a characteristic value for apatites [43], is clearly visible. According to [44], in the case of substitution of phosphate ions by carbonate ions, the phosphate line ν1 appears in the range of 955-959 cm–1. The line of symmetric stretching vibration ν1(PO) was detected at 961 cm–1. This fact indicates the high crystallinity of fluorapatite, which is confirmed by the XRD data (Fig. 4). In addition, the position of the ν3 line (1038 cm–1) corresponding to the asymmetric P-O stretching vibration did not change. For fluorapatite Ca8SrMg(PO4)6F2, a decrease in the intensity of the ν3 line and its fusion with the extended ν1 line are observed (Fig. 8b). In contrast to the ν1 and ν3 lines, the remaining lines of the P-O deformation vibrations ν2 (423→413 cm–1) and ν4 (596→606 cm–1) are shifted towards lower frequencies when calcium is replaced by strontium and magnesium. In addition, on the Raman spectra of fluorapatite and strontium-magnesium-containing fluorapatite, a 1077 cm-1 line was observed. This line is corresponded to CO32- vibrations and indicated partial replacement of phosphate ions by carbonate ions
(CO32- → PO43-). The presence of CO32– ions is caused by the absorption of atmospheric carbon dioxide by synthesized fluorapatite powders. It is known that the PO43- tetrahedron in the fluorapatite lattice is surrounded by 9 calcium cations, which isolate it from other PO43-tetrahedra. In [45] it is reported that the substitution of PO43- ions by CO32– ions does not cause any significant structural changes in the FAp structure that could be observed on the Raman spectra. These results are also in a good agreement with the IR analysis data.
Long-term immersion tests demonstrate the real behavior of the biomaterials in biological environment. In this regard, the solubility of the synthesized fluorapatites in physiological media and the effect of the structural replacement of calcium by strontium and magnesium are of great interest.
The results of the solubility tests of fluorapatites Ca10(PO4)6F2, and Ca8MgSr(PO4)6F2 in saline solution are shown in Fig. 9a. A decrease in the degree of dissolubility of fluorapatite in time was observed. These results also indicate a greater degree of solubility of strontium-magnesium substituted fluorapatite in comparison to unsubstituted fluorapatite. Fig. 9b shows the results of the solubility of the synthesized fluorapatites in buffer solution. It should be mentioned that in the previous study [38] an increase of the weight losses and dissolution rates of strontium substituted fluorapatite Ca9Sr(PO4)6F2 in comparison to unsubstituted fluorapatite Ca10(PO4)6F2 samples was detected. Based on [2], this fact is associated with the presence of tricalcium phosphate Ca3(PO4)2 phase in the Ca9Sr(PO4)6F2 samples. It is known that tricalcium phosphate demonstrate greater solubility in comparison with fluorapatite. In the contrast, it was reported in [46] that the strontium and calcium substitutions cause no principal change in the dissolution rate of phosphate glasses. Additionally, the effect of strontium bonding to a similar number of phosphate chains as calcium was found. The addition of Sr into the phosphate glass composition resulted in a decrease of the dissolution rate of the glass, thus suggesting an increase of the cross-linking between phosphate chains. At present study, the results demonstrate that in both saline and buffer solutions the solubility of strontium-magnesium substituted fluorapatite is higher than fluorapatite. It should be noted that in the buffer solution, the solubility of fluorapatites was not much higher than the solubility in saline.
There is a strong correlation between the structure, mechanical properties and solubility of ceramics. The ceramics with higher density and crystallinity demonstrate better mechanical characteristics and biocompatibility [47]. The hardness (H) to the Young’s modulus (E) ratio is the main parameter, which characterize the material deformation in relation to yielding [48]. H/E ratio plays an important role in identifying the mechanical behaviors and further brittle failure of ceramic materials and coatings [49]. Furthermore, fracture toughness (K1c) parameters demonstrate the ability of a material to resist against brittle failure and crack propagation. Nanoindentation method allows to make direct measuring of cracks created with a sharp diamond indenter [50, 51].
Taking the above-mentioned points into consideration, the main mechanical parameters of Ca8SrMg(PO4)6F2 and Ca10(PO4)6F2 fluorapatites before and after immersion tests are presented in Table 2. The measured average value of Young's modulus on the surface of fluorapatite slightly exceeds the value for strontium-magnesium doped fluorapatite. After a solubility test in saline solution, the values of Young's modulus measured on the surface of both samples slightly decrease. On the contrary, microhardness and K1c parameters for Ca8SrMg(PO4)6F2 samples decreased to a greater extent after immersion tests possibly due to the higher soluble tricalcium phosphate Ca3(PO4)2 phase content.
Table 2. Young's modulus, hardness and fracture toughness parameters of fluorapatite samples before and after dissolution tests
Fluorapatites
|
Modulus, GPa
|
Hardness, GPa
|
K1c, MPa · m1/2
|
Before dissolving
|
After dissolving
|
Before dissolving
|
After dissolving
|
Before dissolving
|
After dissolving
|
Ca10(PO4)6F2
|
106.3±4.5
|
97.7±6.1
|
5.9±1.1
|
5.2±0.6
|
2.0±0.9
|
1.7±0.5
|
Ca8MgSr(PO4)6F2
|
102.7±5.3
|
93.4±4.7
|
5.5±0.7
|
4.1±1.9
|
1.8±0.4
|
1.3±0.8
|
Nevertheless, the dissolution of the synthesized fluorapatites within 14 days occurs to a small extent and does not significantly affect their mechanical properties.
The surface of fluorapatite samples after 14 days of immersion tests in saline solution was studied by Raman spectroscopy, SEM, and laser confocal microscopy.
An analysis of the Raman spectra of the surface of fluorapatite samples after solubility tests also showed the presence of the main fluorapatite lines ν1(P-O) at 965 cm-1, see Fig. 10. These results indicate a strong stability of fluorapatite and strontium-magnesium-doped fluorapatite in saline solution at a temperature of 37 ºC. The presence of the ν1 peak in the Raman spectrum confirms the stability of the crystal structure of fluorapatite on the surface of the studied samples. After solubility tests for Ca10(PO4)6F2 fluorapatite, a shift to the region of high frequencies of the line ν3 (1038→1052 cm-1) and low frequencies of the line ν2 (423→430 cm-1) were observed in (Fig. 10a). On the contrary, the line ν4 was shifted to the low frequencies region
(596 → 588 cm-1). At the same time, the position of the line corresponded to the vibrations of CO32- (1077→1079 cm-1) was almost unchanged. For Ca8MgSr(PO4)6F2 fluorapatite after immersion tests, shift of the lines of P-O deformation vibration ν2 (413→425 cm-1) and ν4 (606→628 cm-1) towards higher frequencies was detected (Fig. 10b).
The position of the line corresponding to the vibrations of CO32- was also significantly changed (1077→1088 cm-1). It is known that the solubility of apatite increases as a result of the replacing the PO43– ions with the CO32– ions. The higher solubility of carbonate-containing apatite compared to carbonate-free apatite is partially due to the fact that Ca2+-CO32– bonds are weaker than Ca2+-PO43– bonds [52].
The presence of a second, more soluble, Ca3(PO4)2 phase and partial replacement of phosphorus ions by carbon ions led to increasing in solubility level and, accordingly, to changes in the surface of fluorapatite samples after dissolution, which are reflected in the Raman spectra.
The surfaces of the Ca10(PO4)6F2 and Ca8MgSr(PO4)6F2 samples after immersion tests, are significantly different, see Fig. 11a and Fig. 11b. It is noticeable that the surface of the Ca8MgSr(PO4)6F2 sample underwent stronger dissolution as compared to Ca10(PO4)6F2, see Fig. 11c and Fig. 11d.
According to the data of EDX analysis of surface after soaking tests, the main peaks of Ca, P, O, F, Sr and Mg corresponding to the elemental composition of fluorapatites are identified in Fig.12.
The EDX spectra demonstrate some changes in the peak intensity after immersion tests. A noticeable decrease of the fluorine content in the composition of the fluorapatites samples was detected. C. Chaïrat et al [53] argue that the dissolution of apatite is primarily due to the relatively rapid removal of F and Ca from the contact surface. Thus, the destruction of fluorapatite occurs due to the breaking of Ca–O bonds on the surface depleted of calcium and fluorine.
The stability of structure, phase composition, and mechanical properties during long term period of staying in biological environment is very important for biocompatibility of ceramics, if they are used as a biomaterial. Thus, both the Ca10(PO4)6F2 and Ca8MgSr(PO4)6F2 fluorapatites can be further proposed as promising candidates for biomedical use in the replacement of defective areas of bone.