Next Article in Journal
Powders Based on Ca2P2O7-CaCO3-H2O System as Model Objects for the Development of Bioceramics
Next Article in Special Issue
Effects of the Processing Technology of CVD-ZnSe, Cr2+:ZnSe, and Fe2+:ZnSe Polycrystalline Optical Elements on the Damage Threshold Induced by a Repetitively Pulsed Laser at 2.1 µm
Previous Article in Journal
Oxidation Resistance of γ-TiAl Based Alloys Modified by C, Si and Y2O3 Microdopants
Previous Article in Special Issue
Thermoelectric Properties of Si-Doped In2Se3 Polycrystalline Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Possibilities of Mechanochemical Synthesis of Apatites with Different Ca/P Ratios

by
Marina V. Chaikina
1,
Natalia V. Bulina
1,*,
Olga B. Vinokurova
1,
Konstantin B. Gerasimov
1,
Igor Yu. Prosanov
1,
Nikolay B. Kompankov
2,
Olga B. Lapina
3,
Evgeniy S. Papulovskiy
3,
Arcady V. Ishchenko
3 and
Svetlana V. Makarova
1
1
Institute of Solid State Chemistry and Mechanochemistry, Siberian Branch of the Russian Academy of Sciences, Kutateladze Str. 18, 630128 Novosibirsk, Russia
2
A.V. Nikolaev Institute of Inorganic Chemistry, Siberian Branch of Russian Academy of Sciences, Lavrentyev Ave. 3, 630090 Novosibirsk, Russia
3
G.K. Boreskov Institute of Catalysis, Siberian Branch of Russian Academy of Sciences, Pr. Akad. Lavrentieva 5, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Ceramics 2022, 5(3), 404-422; https://doi.org/10.3390/ceramics5030031
Submission received: 15 July 2022 / Revised: 29 July 2022 / Accepted: 30 July 2022 / Published: 3 August 2022
(This article belongs to the Special Issue Advances in Ceramics)

Abstract

:
Apatite is widely used in medicine as a biomaterial for bone tissue restoration. Properties of apatite depend on its composition, including the Ca/P ratio. This paper shows what range of Ca/P ratio can be attained in apatite by the mechanochemical method of synthesis, providing fast formation of a single-phase product. The synthesis was carried out from a reaction mixture of CaHPO4 and CaO at different Ca/P ratios in the range of 1.17–2.10. The products were studied by PXRD, FTIR and NMR spectroscopy, HRTEM, and STA. In mixtures with a low initial Ca/P ratio (1.17–1.48), directly in the mill, the formation of calcium orthophosphate with whitlockite structure containing an HPO42− group and structural water is shown for the first time. This phosphate has structure similar to that of whitlockites of hydrothermal origin and differs from high-temperature β-tricalcium phosphate that has composition Ca3(PO4)3. A series of samples of apatite was obtained with varied composition, which depends on the initial Ca/P ratio. At Ca/P < 1.67, the formation of two types of calcium-deficient apatite was documented. At Ca/P > 1.67, the existence of two types of calcium-rich apatite is confirmed.

1. Introduction

According to extensive literature, the prototype of the mineral component of bone and dental tissues in humans and animals is hydroxyapatite, Ca10(PO4)6(OH)2 (HAp), modified with various ions at different lattice positions [1,2,3,4]. Bioapatite from hard tissues are usually poorly crystallized apatite nanocrystals [5,6] sometimes with an amorphous phase [7]. The similarity of the composition of HAp with that of the mineral component of bone has led to widespread use of synthetic HAp in medicine as a filler material for the repair of bone defects, as drug delivery systems, as an additive in toothpastes and detergents, for the production of ceramic implants or coatings on metal implants [8,9], and for the preparation of bone cements [10]. The β-form of tricalcium phosphate Ca3(PO4)2 (β-TCP) having the Ca/P ratio of 1.5 is also widely employed as a material in clinical practice [11]. Sometimes, it is used in the form of a two-phase mixture with HAp, as a more easily soluble and bioavailable material [12]. Octacalcium phosphate Ca8H2(PO4)6·5H2O (OCP) having lower Ca/P ratio, namely 1.33, is also used in clinical practice for implantation into bone defects [8].
The ratio of calcium atoms to phosphorus in stoichiometric HAp is 1.67. Bioapatites are characterized by a deficiency of calcium ions and wide variation of the Ca/P ratio, which is sometimes much lower than that of stoichiometric HAp. Such apatites are usually called calcium-deficient apatite (CDAp) [1]. Compounds with Ca/P > 1.67 having apatite crystal structure were named calcium-rich apatites (CRIAp) [1]. Investigations into properties of synthetic apatites with different Ca/P ratios have shown that this ratio is an important parameter that affects the properties of the material, in particular its solubility and thermal stability [8], which is crucial for practical application.
Great interest in the composition and mechanism of formation of nonstoichiometric apatites is associated with a major problem: processes of calcification in various vitally important organs, painful deposits on them in the form of nonstoichiometric compounds, and changes in the composition of bone tissues depending on various conditions and causes. Previously, it has been thought that biominerals—namely, calcified deposits on heart valves and vessels of various organs—are generated by precipitation from the surrounding fluid [1]. After that, first, a model was proposed [13], and later in vivo research revealed that apatite biominerals are formed by a precursor mechanism, where OCP precipitates first, and then it is reversibly hydrolyzed in situ to an intermediate product of the transition from OCP to HAp [14].
A substantial number of studies have been devoted to the synthesis of CDAp at various Ca/P ratios [11,12,15,16,17,18]. The authors of refs. [16,17] have synthesized CDAps with Ca/P ratio within 1.5 ≤ Ca/P ≤ 1.67 by precipitation and hydrolysis methods, respectively. The authors of paper [18] showed that hydrolysis can be used for the synthesis of CDAp with a Ca/P ratio below 1.5. Hydrolysis of a reaction product of a mixture of solutions of calcium nitrate and ammonium phosphate in the presence of hexamethylenetetramine results in the composition of the final product corresponding to the formula Ca8.5(HPO4)2(PO4)4OH·H2O, i.e., Ca/P = 1.416. Such a Ca/P ratio is observed in amorphous biological phosphates [1]. More “acidic” phosphates reaching a value of 0.5 have been synthesized by chemical precipitation in ref. [19]. In the samples with Ca/P ratios of 2.0 and 2.5, those authors registered CaO in addition to the dominant phase of HAp, indicating the impossibility of CRIAp formation by the precipitation technique.
Single-phase CRIAp samples with Ca/P = 1.73 have been obtained by prolonged heating of stoichiometric HAp with CaCO3 in saturated water vapor at 1000 °C [20]. A series of CRIAp with Ca/P = 1.67–1.9 has been synthesized from a suspension of calcium hydroxide and a solution of phosphoric acid at pH 7.5–11.5 [21]. While studying thermal decomposition of CDAp by transmission electron microscopy (TEM), the authors of ref. [22] discovered a metastable phase arising in a narrow temperature range (700–800 °C). According to energy-dispersive X-ray analysis, this metastable phase has a higher Ca/P ratio (>1.67) than that of the HAp matrix.
We did not find research articles on the synthesis of CDAp or CRIAp by the mechanochemical method, which, under certain conditions, is a simple and fast way to obtain HAp [23]. In the mechanochemical synthesis, chemical reactions are initiated by the energy released during ball collisions and due to the action of friction forces. The released energy depends on technical characteristics of the mill, namely, kinetic energy of the balls (vial rotation speed). In this regard, the synthesis of HAp occurring directly in the milling vials is possible only in ball mills working at a high rotation speed. The higher the rotation speed of the vial, the shorter is the time needed for the HAp synthesis. The aim of the present work was to study the possible range of the Ca/P ratio in HAp forming during mechanochemical synthesis in a high-energy planetary ball mill at a rotation speed of 1800 rpm.

2. Materials and Methods

The initial components for the synthesis of samples were monetite CaHPO4 (of pure grade, Vekton) and freshly calcined at 1000 °C calcium oxide (CaO; analytical grade, Vekton). The reagents were mixed in different ratios so that the Ca/P ratio varied from 1.17 to 2.1 (Table 1). Monetite was chosen as the initial calcium phosphate because the use of other calcium phosphates, such as Ca(H2PO4)2·H2O and CaHPO4·2H2O, is accompanied by cementation and aggregation of the mixture during activation [23]. The synthesis of the sample with Ca/P = 1.33 suggested the formation of OCP having five water molecules. In this regard, the missing amount of distilled water was added to the initial reaction mixture (Table 1). Since there are no additional procedures in the synthesis other than mixing, it is obvious that all introduced elements are present in the sample, which means that the Ca/P ratio is preserved. In this regard, an elemental analysis of synthesized samples was not carried out.
The mechanochemical synthesis was conducted in a planetary ball mill of the AGO-2 type (ISSCM SB RAS, Russia) [24] with water-cooled vials 150 mL each at a rotation speed of 1800 rpm. The weight of the milling balls was 200 g. The weight ratio of the reaction mixture to the milling balls was 1:20. Based on ref. [23], the duration of the mechanochemical treatment was 30 min, which is sufficient for complete conversion of the initial reagents into the HAp phase. To avoid contamination of the product by the material of the balls and vials, the working zone of the vial was lined with the reaction mixture beforehand.
Products of the mechanochemical synthesis were assessed by powder X-ray diffraction (PXRD) analysis, Fourier transform infrared (FTIR) spectroscopy, TEM, nuclear magnetic resonance (NMR) spectroscopy, and simultaneous thermal analysis (STA).
The PXRD patterns were recorded on a D8 Advance powder diffractometer (Bruker, Karlsruhe, Germany) in Bragg–Brentano geometry using CuKα radiation, a nickel Kβ-filter, and an ultrafast position-sensitive one-dimensional Lynx-Eye detector. The phase analysis of the compounds was carried out in the ICDD PDF-4 database (a 2011 release). The unit cell parameters, crystallite size, and the phase concentrations were determined after computation of background coefficients and sample displacement refinement by the Rietveld method in Topas 4.2 software (Bruker, Germany). The instrumental contribution was calculated by the method of fundamental parameters. The average size of the crystallites was estimated using the Lorentz convolution, which varies in 2Θ as a function of 1/cos (Θ). The Chebyshev polynomial of the eighth order and the 1/x function were employed to describe the background.
FTIR spectra were recorded using an Infralum-801 device (Simex, Novosibirsk, Russia). The samples were prepared by the traditional KBr pellet method.
TEM and high-resolution TEM (HRTEM) images were obtained using a Themis-Z 3.1 microscope (TFS, Waltham, MA, USA) at an accelerating voltage of 200 kV. The microscope is equipped with a field emission cathode having a monochromator and with two aberration correctors. Energy-dispersive X-ray microanalysis of elemental composition of the samples was performed on a four-segment Super-X detector (with an energy resolution of ~120 eV) in scanning dark-field mode with the construction of maps of distributions of elements by means of characteristic lines of the spectrum from each point in an analysis region. Samples for the electron-microscopic examination were dispersed by ultrasonication and deposited from alcohol on a substrate: copper perforated grids 3 mm in diameter covered with a thin carbon mesh.
Solid-state NMR experiments were conducted on a Avance III 500 spectrometer (Bruker, Karlsruhe, Germany) at a resonance frequency of 500 MHz for 1H experiments and 200 MHz for 31P experiments. Magic angle spinning (MAS) spectra were acquired using a 4 mm probe at sample rotation frequency 10 kHz.
STA included simultaneous detection of mass loss, differential scanning calorimetry (DSC), differential thermal analysis (DTA), and registration of the evolved gas using a mass spectrometer. STA was carried out by means of a STA 449 F1 Jupiter device (Netzsch, Lesb, Germany) equipped with a QMS 403 C Aeolos mass spectrometer. The measurements were performed in a Pt–10wt%Rh crucible under an argon-oxygen mixture (80:20) at a heating rate of 10 °C/min.

3. Results

3.1. PXRD

The mechanochemical synthesis of a series of samples with different Ca/P ratios was carried out in accordance with Table 1. Figure 1 shows selected PXRD patterns of the synthesized samples that have obvious differences from each other. PXRD patterns of the other samples match in appearance those presented in Figure 1. All the PXRD patterns contain clear-cut reflections indicating the presence of crystalline phases. Figure 1 shows that reflections of an HAp phase are absent in samples with Ca/P < 1.25. Nevertheless, the modeling of PXRD patterns by the Rietveld method revealed that in the region of the main reflection of HAp (2θ 31.7°), there remains a certain background region, which we assigned to semiamorphous apatite. When the diffraction patterns were modeled taking into account the HAp phase with a crystallite size of less than 18 nm, reliability factors improved considerably. Given the very low degree of crystallinity of this phase, it is named as an amorphous apatite in Table 2.
The results of phase composition analysis of the samples, as performed by the modeling of phases by the Rietveld method, are given in Table 2. This technique for analyzing diffraction patterns allows to refine crystal lattice parameters of the phases and their crystallite size, to calculate the number of phases, and to detect a minor amorphous phase. Table 2 shows that there are three Ca/P ratio ranges: single-phase (Ca/P = 1.50–1.90), two-phase (Ca/P = 1.33–1.48, 2.00), and three-phase (Ca/P = 1.25–1.17, 2.10). In the single-phase range, Ca/P = 1.5–1.9, only HAp is present. In samples with Ca/P = 1.33–1.48, in addition to HAp, a phase corresponding to β-TCP structure was registered. Reflections of this phase match card No. 040-14-2292 from the ICDD PDF-4 database. At Ca/P ≤ 1.25, the phase of the initial unreacted monetite (card No. 010-71-1759) is seen in addition to the above. In these samples, the HAp phase has a nanoscale amorphous state and a low concentration. The crystallite size is estimated to be 10–15 nm. The diffraction pattern does not contain clear-cut reflections of HAp (Figure 1), and therefore it is difficult to determine lattice parameters of this phase. At Ca/P = 2.0, in addition to HAp (card No. 010-71-1759), a phase of the initial CaO (card No. 010-82-1691) is present in the sample, and at Ca/P = 2.1, aside from the above, a Ca(OH)2 phase (card No. 000-44-1481) is seen.
Figure 2a,b shows the dynamics of changes in apatite crystal lattice parameters with the increasing initial Ca/P ratio. As Ca/P goes up from 1.33 to 1.90, parameter a gradually decreases and stays virtually unchanged in the range of 1.9–2.10. The c parameter shows a different dependence, namely, it diminishes up to the ratio Ca/P = 1.5 and then begins to grow. It is noteworthy that the boundaries of the single-phase range coincide with a shift in the nature of changes in the lattice parameters: at the lower limit, dynamics of parameter c change, and at the upper limit, there are changes in the dynamics of parameter a (Figure 2a,b). The size of crystallites in the single-phase range varies between 20 and 21 nm (Figure 2c), whereas for 1.5 > Ca/P > 1.9, a decrease in this size is observed.
Lattice parameters of the β-TCP phase in the range Ca/P = 1.17–1.38 do not show any pronounced dynamics (Figure 3). With a further increase in Ca/P, parameter a begins to rise while c starts to decline. It should be noted that the observed parameters’ values differ substantially from those of high-temperature β-TCP that were obtained by the classic ceramic method; in the latter case, the material has parameters a = 10.427(1) Å and c = 37.456(4) Å [25].

3.2. FTIR Spectroscopy

Figure 4 presents FTIR spectra of synthesis products from mixtures having different initial Ca/P ratios. According to the nature of the spectrum, the data were subdivided into four ranges of changing composition of the samples. It is more convenient to start the analysis of these spectra with the identification of the spectrum of stoichiometric HAp having Ca/P = 1.67 (Figure 4d). This FTIR spectrum contains all absorption bands characteristic of stoichiometric HAP: bending ν4 vibrations of O–P–O bonds (570 and 603 cm1), stretching vibrations ν1 (960 cm1) and ν3 (1046 and 1088 cm1) of the P–O bond, and libration (630 cm1) and stretching vibrations (3572 cm1) of the OH group in HAp (Figure 4d) [1]. There are weak absorption bands at 870, 1416, and 1493 cm1 due to bending and stretching vibrations of C–O bonds of carbonate groups in apatite [1]. Mechanochemical synthesis of HAp in ambient air is usually accompanied by the incorporation of a small number of CO32 groups into the lattice, and this number depends on the synthesis conditions [1].
In the spectra of samples with Ca/P > 1.67, positions of the absorption bands of HAp remain the same, and a small number of carbonate groups is present everywhere (Figure 4d). For samples with Ca/P = 1.9 and higher, an additional band at 3640 cm1 appears in the spectrum, and this band’s intensity grows with increasing Ca/P (Figure 4d). This band belongs to vibrations of bonds of the OH groups from calcium hydroxide Ca(OH)2. It should be mentioned that according to our diffraction data, Ca(OH)2 is detectable only in the sample with the highest Ca/P ratio (Table 2). In this case, infrared spectroscopy turned out to be the more sensitive method.
In samples with Ca/P from 1.60 to 1.50 (Figure 4c), i.e., lower than that of stoichiometric HAp, there are no absorption bands of C–O bonds of carbonate groups. Nonetheless, in addition to the absorption bands of HAp, a wide low-intensity absorption band is seen at 872 cm1 (Figure 4c), indicating the presence of HPO42 groups in the apatite lattice [1].
In the range 1.33 ≤ Ca/P ≤ 1.48, the band of HPO42 groups is still present (Figure 4b). A shift of absorption bands of the phosphate group to a lower-frequency region is observed, as is a change in the ratio of their intensities. The latter effect may be due to the presence of absorption bands of an additional phase (β-TCP), as revealed by PXRD data (Table 2). Moreover, with a decrease in Ca/P, the absorption band of the stretching vibration of the HAp hydroxyl group at 3570 cm−1 disappears in the spectra, thus pointing to a decline of the number of hydroxyl groups.
Samples with 1.17 ≤ Ca/P ≤ 1.25 have a spectrum that is much different from that of HAp (Figure 4a). Obvious absorption bands of HAp are absent in these FTIR spectra, and hence HAp is absent in the crystalline state. According to X-ray phase analysis (Table 2), aside from amorphous HAp, these samples contain β-TCP and CaHPO4. For comparison, Figure 4a shows FTIR spectra of these substances in their pure form. Readers can see that the spectra of the samples with Ca/P = 1.17 and 1.25 are not the sum of the spectra of CaHPO4 and β-TCP. It can be hypothesized that these are precursors of “acidic” orthophosphates along with fragments of the structure destroyed during activation. The spectra of the analyzed samples are close in shape to the β-TCP spectrum; moreover, they have two bands, at 553 and 606 cm1, matching in position. The rest of the bands are considerably shifted, implying an alteration in the immediate environment of β-TCP’s phosphate tetrahedra. Additionally, in samples with Ca/P = 1.17–1.25, there is a band at 888 cm1 from the HPO42 group, which was observed at 872 cm1 in the samples with high Ca/P (Figure 4b) and is present in the spectrum of CaHPO4 at 894 cm1 (Figure 4a). The absorption band at 1645 cm1 belongs to adsorbed water. It must be noted that the FTIR spectra of samples with Ca/P = 1.17–1.25 are also similar to the spectra of amorphous phosphates [1,7].

3.3. TEM

Figure 5 depicts electron micrographs of samples with different Ca/P ratios. All the samples are aggregates of nanoparticles ranging in size from 10 to 1000 nm. X-ray microanalysis data of the samples are shown in Figure 5. According to the results of energy-dispersive X-ray spectroscopy, in samples with low Ca/P, the distribution of elements among particles is uniform, whereas in the sample with the highest Ca/P value, different concentrations of elements are observed among particles, but the average is 2.00 (Figure 6c). The deviation of the observed Ca/P ratio from the intended ratio in the samples with Ca/P = 1.25 and 1.42 can be explained by the finding that these samples contain several phases with different Ca/P ratios, including unreacted monetite, where Ca/P = 1.00.
It is worth mentioning that in the samples with Ca/P = 1.25 and 1.42, some particles contained spherical cavities in the form of bubbles. Under the influence of the electron beam, such particles quickly transformed into monolithic ones (Figure 7, a link to a video is given in Supplementary Information).

3.4. NMR

31P and 1H MAS NMR spectra of some mechanochemically synthesized samples are given in Figure 8. The NMR parameters of the observed lines are given in Table 3 and Table 4. The line fitting was carried out in accordance with literature data [26,27,28,29,30,31,32].
The signal of phosphate tetrahedra of the HAp phase (δiso ≈ 2.9 ppm) is not observed in the 31P NMR spectrum of the sample having Ca/P = 1.25. The spectrum contains two sets of lines: a broad base with a maximum at 1.2 ppm (a superposition of at least two signals with a chemical shift of 2.5 and 0 ppm) and a broad low-intensity line at –8 ppm (a superposition of at least three signals with a shift of −7, −9, and −10 ppm). According to the literature data, the peaks at −7.43, −9.09, and −10.35 ppm indicate the presence of a low amount of calcium pyrophosphate Ca2P2O7 [26,27]. The main phase in the sample with Ca/P = 1.25 is shown in the 31P MAS NMR spectra as a broadened signal with a maximum at 1.2 ppm, which cannot be ascribed to β-TCP, as it could be proposed according to PXRD data. The 31P MAS NMR spectrum of β-TCP is extremely specific and is described by several of lines, that can be categorized into two main sets: at 5 and 0 ppm [26,27,28,29], both are absent in the sample having Ca/P = 1.25 sample. The modification of β-TCP with a substitution of calcium with zinc or strontium leads to a major change in the spectrum; however, this kind of signal is absent [29]. A similar change in the 31P MAS NMR spectrum of β-TCP has been observed in the composite “β-TCP–inositol phosphate (IP6),” but the line is broader in the spectrum of this composite than in our case [31]. This substantial change in structure of β-TCP structure is likely to be caused by the presence of HPO42− and a large amount of retained water exchanging with the HPO42− group, as deduced from the 1H MAS NMR spectrum of the sample having Ca/P = 1.25 (Figure 8b). Interaction of molecular water with orthophosphate groups in amorphous calcium phosphates also leads to considerable broadening of the 31P spectrum [33]. The signals of HPO42− groups observed in the 1H NMR spectrum at 16.00 and 12.26 ppm can be attributed to the CaHPO4 phase [32].
A signal from phosphate tetrahedra in apatite dominates in the 31P NMR spectra of the samples with Ca/P > 1.25 (Table 3). The narrowest signal with isotropic shift of 2.85 ppm seen in the sample with Ca/P = 1.67 correspond to the structure of stoichiometric HAp. In the samples with a large and lower amount of Ca, the signal of phosphate tetrahedra in apatite is broader and shifted to low field. There is also a broadening of the signal of a hydroxyl group in the 1H spectra (Table 4).
It should be noted that the samples with Ca/P = 1.50 and 1.67 contains a greater amount of retained water (~50%; Table 4). In the sample having Ca/P = 2.00, the content of fixed water is lower: 30%; however, in the 1H spectrum, the signal at 1.63 ppm—corresponding to a hydrogen atom bonded to the hydroxyl group of HAp (Table 4)—is stronger.

3.5. STA

Figure 9 presents STA data for samples with Ca/P ≤ 1.67 and illustrates how the DSC curves become more complicated as the Ca/P ratio of the initial mixtures decreases. This finding indicates that the composition of the obtained products becomes more complicated too. In all samples, a mass loss is observed owing to the evaporation of adsorbed water, with a maximum endothermic effect between 100 and 120 °C (Figure 9a–f). Weakly bound lattice water is released at 200 °C. With a further increase in temperature during the acquisition of STA data, changes in the mass loss depend on the composition of each sample and on thermal stability of the synthesized products. Additional endothermic effects manifest themselves at lower temperatures, namely at 365–370, 430, and 650–600 °C (Figure 9a,b). These endothermic effects are due to the decomposition of unreacted monetite and of precursors of “acidic” orthophosphates with fragments of the destroyed structure during the mechanochemical synthesis. The decomposition gives rise to pyrophosphate and water.
In the sample with the Ca/P ratio of 1.67, corresponding to stoichiometric HAp, the release of adsorbed water (~2%) gradually turns into a lattice water release, which lasts up to 500 °C and raises the mass loss to 3.8% (Figure 9f). The further water release observed at >700 °C is explained by HAp dehydroxylation [34].
At Ca/P = 1.50 (Figure 9e), aside from the above phenomena, an additional process takes place at 739 °C and corresponds to the decomposition of CDAp with a release of β-TCP, HAp, and water (Table 5 and Figure 9e) according to the reaction:
Ca10x(HPO4)x(PO4)6x(OH)2−x → (3x)Ca3(PO4)2 + (1 − x)Ca10(PO4)6(OH)2 + H2O + (1 − x)O2,
where x ≤ 1.
After STA, the sample with Ca/P = 1.50 was biphasic (Table 5). It consisted of 92 wt% of β-TCP and 8 wt% of HAp. FTIR data confirmed the predominant formation of β-TCP in this sample (Figure 10). The spectrum shows all characteristic absorption bands consistent with Figure 3a, which presents the spectrum of a β-TCP standard.
The behavior of STA curves of the sample with Ca/P = 1.42 differs from that of the samples above by the presence of an additional endothermic effect at 205 °C because of a more intense release of lattice water (Figure 9d). After STA, in the sample with Ca/P = 1.42, HAp was no longer detectable, there was 90 wt% of β-TCP and 10 wt% of Ca2P2O7 (Table 5). The pyrophosphate is a decomposition product of more “acidic” CDAp presented in the sample with Ca/P = 1.42 as compared to the sample with Ca/P = 1.50.
In the STA curves of the sample with Ca/P = 1.33 (Figure 9c), one more endothermic process is present with a water release at 638 °C. It is worth noting that the same phases are present in this material after STA as in the sample with Ca/P = 1.42, where this endothermic effect is absent. We believe that the endothermic process at ~640 °C is also attributable to the release of β-TCP, but at a lower temperature. In the sample with Ca/P = 1.33, there are probably two types of β-TCP precursors that decompose with the release of water at ~640 and 750 °C.
In the samples with Ca/P = 1.25 and 1.21, the water release curve clearly indicates that the observed additional endothermic effects at ~370 and ~430 °C are related to the release of water during the decomposition of unreacted monetite and of precursors of “acidic” orthophosphates with fragments of the structure destroyed during the mechanochemical synthesis (Figure 9a,b). It should also be noted that in the samples with Ca/P = 1.25 and 1.21, the formation of the β-TCP phase occurs in one stage and much earlier than in the above-mentioned samples (at 604–625 °C), suggesting that the composition of the β-TCP precursor changes. In the samples after STA, only β-TCP and Ca2P2O7 were present, as in the samples above (Table 5).

4. Discussion

In this work, reactions of the synthesis of phosphates—from initial components CaHPO4 and CaO at different ratios—cover the range from Ca/P of 1.17 to 2.10. This wide range includes (i) low Ca/P ratios characteristic of CDAp, (ii) stoichiometric HAp having Ca/P = 1.67, and (iii) the range of CRIAp up to Ca/P = 2.10 (Table 1). Among these values, there are Ca/P ratios belonging to other well-studied calcium orthophosphates such as OCP (Ca/P = 1.33) and TCP (Ca/P = 1.5).
According to the obtained analytical data, the studied Ca/P range can be divided into several subranges: (1) Ca/P ≤ 1.25, (2) 1.33 ≤ Ca/P ≤ 1.48, (3) 1.5 ≤ Ca/P ≤ 1.60, (4) 1.67 ≤ Ca/P ≤ 1.90, and (5) Ca/P ≥ 2. Each subrange is characterized by its own structural features.

4.1. Range Ca/P ≤ 1.25

In the Ca/P ≤ 1.25 range, synthesis products contain unreacted monetite, amorphous apatite (most likely nuclei of CDAp and pyrophosphate), and a phase with β-TCP whitlockite structure; however, its lattice parameters differ considerably from those of high-temperature β-TCP as described above. Our FTIR and NMR spectroscopy findings also indicate some distinctive features of the β-TCP obtained in the current experiments. It possesses an HPO42 group and a large amount of structural water, which is in exchange with the proton of the HPO42 group. For this reason, samples with Ca/P ≤ 1.25 may begin to “come to life” and crystallize under the action of an electron beam of a microscope. The highest concentration of the whitlockite-like phase is observed in samples with Ca/P = 1.21 and 1.25 (Table 2). A further decrease or increase in Ca/P lowers its concentration. The formation of calcium phosphates with whitlockite structure but with composition more “acidic” than that of β-TCP is quite possible, because in our case, an excess of CaHPO4 is observed in the initial mixture.
The structure similar to whitlockite β-TCP and containing an “acidic” group is present in a variety of natural and synthetic compounds. Such natural minerals as calcium whitlockite Ca18.19(Mg1.17Fe0.63)H1.62(PO4)14 [35] and strontium whitlockite Sr9Mg(PO3OH)(PO4)6 [36] have been found. Calcification sites in various human and animal organs also have whitlockite structure [35]. Proton-bearing whitlockites Ca18Mn2H2(PO4)14 [37], Ca18Mg2H2(PO4)14 [35], and Ca9(Fe0.63Mg0.37)H0.37(PO4)7 [38] have been synthesized by the hydrothermal method. All the compounds listed in this paragraph have a PXRD pattern identical to that of high-temperature whitlockite β-TCP and a Ca/P ratio of <1.5.
It can be theorized that under our conditions, a mixture of CaO and CaHPO4 with an excess of the latter forms proton-containing whitlockite with the possible chemical formula Ca10(PO3OH)(PO4)6. The interaction of the reagents involves several stages featuring intermediate products–precursors, and then the proton-containing whitlockite comes into being:
CaHPO4 + Ca(OH)2 → Ca2(PO4)OH (precursor) + H2O
Ca2(PO4)OH (precursor) + CaHPO4 → Ca3(PO4)2·H2O
3Ca3(PO4)2·H2O + CaHPO4 → Ca10HPO4(PO4)6·3H2O
The formula Ca10(HPO4)(PO4)6 is similar to that of the CDAp that is devoid of a hydroxyl group. The Ca/P ratio in this compound is 1.428, which can be considered as the highest possible substitution of calcium with a proton and the limit of the existence of CDAp transitioning into proton-containing whitlockite.
It was quite unexpected for us to obtain whitlockite-like structure by the mechanochemical method directly in the mill. It is known that β-TCP is a high-temperature form of calcium phosphate and emerges after heat treatment of a mixture of reagents with Ca/P = 1.50 at 650–750 °C [39,40]. A question arises: how can a high-temperature compound form during mechanochemical synthesis at the moderate temperature? According to literature data [41], under certain conditions of mechanochemical synthesis in powerful planetary ball mills, locally, at the contact point of the balls, the temperature can go up to 600 °C and higher. It is also known that during mechanochemical synthesis, conditions may arise where interactions of the components in the reaction mixture resemble hydrothermal ones [42], and not only high temperature but also high pressure is possible. In that detailed review, the author explains by which processes the high temperature and pressure necessary for the initiation of hydrothermal processes can be reached under the action of mechanical impulses arising in powerful planetary ball mills in solid matter–water systems. Furthermore, experimental data are presented where using several silicates as an example, products of mechanochemical synthesis have been compared with the products obtained in autoclaves. It was demonstrated above that proton-containing whitlockites can form under the conditions of hydrothermal synthesis; therefore, they can also form during mechanochemical synthesis. It is possible that it is hydrothermal processes that underlie the formation of the compound with whitlockite structure during the mechanochemical synthesis of nonstoichiometric calcium phosphates from mixtures having a low initial Ca/P ratio. The presence of HPO42 groups and of the considerable amount of water in the lattice of these calcium phosphates testifies in favor of the hydrothermal genesis of the detected phase possessing whitlockite structure.

4.2. Range 1.33 ≤ Ca/P ≤ 1.48

With a further increase in the Ca/P ratio, starting reagent CaHPO4 is no longer observed among the synthesis products. In the range 1.33 ≤ Ca/P ≤ 1.48, the whitlockite-like phase is present, and distinct reflections of an apatite phase appear (Figure 1). As Ca/P goes up, the concentration of the whitlockite phase declines (Table 2), while its crystal lattice parameters begin to deviate even more from those of stoichiometric β-TCP (Figure 3). It is not possible to elucidate the structure of the whitlockite formed under our conditions of mechanochemical synthesis because it is not monophasic, and its concentration diminishes with increasing Ca/P. Crystal lattice parameters of the apatite phase change too (Figure 2). A decrease in these parameters is seen, accompanied by a strong rise of this phase’s concentration and by crystallite growth (Figure 2, Table 2). The presence of an absorption band of the HPO42 group in the FTIR spectra suggests that these are CDAps. STA findings indicate that they contain a large amount of structural water (Figure 9c,d).
There is no evidence of the presence of OCP having a ratio of Ca/P = 1.33 in the sample with the corresponding Ca/P; thus, it can be concluded that OCP is not formed by this method of mechanochemical synthesis.

4.3. Range 1.50 ≤ Ca/P ≤ 1.60

In this range, only the apatite phase is present among the products of the mechanochemical synthesis (Table 2), which is CDAp. The presence of the HPO42 group is confirmed by FTIR spectroscopy. The ratio of adsorbed to lattice water increases (Figure 9f). Signals from phosphate tetrahedra and hydroxyl groups in apatites are visible in the 31P and 1H NMR spectra (Table 3 and Table 4). Judging by the change in lattice parameters, the samples of CDAp in this range of Ca/P ratios have varied composition approaching stoichiometric HAp. Furthermore, this composition differs from that of samples with Ca/P < 1.5 because the dynamics of parameter c change their direction in this range (Figure 2b). Possible composition of the CDAp within this range is Ca10−x(HPO4)x(PO4)6−x(OH)2−x, where x ≤ 1. The calcium deficiency is compensated by the hydrogen from the HPO42 group, and the electroneutrality of the compound is ensured by vacancies of OH groups.
It should be pointed out that β-TCP, which has a Ca/P value of 1.5, was not found in the sample with the above-mentioned Ca/P ratio. In our case, the formation of proton-containing β-TCP was documented in mixtures with a Ca/P ratio much lower than that of β-TCP. At Ca/P = 1.5, only CDAp is observed as the mechanochemical-synthesis product, which blocks the formation of β-TCP. This result is explained by the finding that the formation of apatite structure is thermodynamically more favorable than the formation of β-TCP: Gibbs free energy (ΔGCDAp) of the formation of CDAp is −11 980 kJ/mol, whereas for β-TCP, ΔGβ-TCP = −3889 kJ/mol. The composition of CDAp at Ca/P = 1.5 matches the formula Ca9HPO4(PO4)5(OH).
With a further decrease in the concentration of the calcium cation, i.e., with a decrease in the Ca/P ratio in apatite, the number of OH groups will be less than one (in terms of electroneutrality). In this context, the crystal lattice of CDAp will contain unit cells in which only OH group vacancies (an empty hydroxyl channel) are contained on the c-axis. The local excess of the positive charge should lead to a rearrangement of the structure into a “channel-less” variant; evidently for this reason, in samples with Ca/P < 1.5, a second phase separates out, which has a lower Ca/P ratio than that of Ca9HPO4(PO4)5(OH) (Table 2). A similar situation of forced phase transformations in the presence of more than one vacancy per unit cell has been reported for silicate-substituted apatites, where the excess charge of the silicate anion is compensated by the emergence of a vacancy of the hydroxyl group [43]. On the other hand, a high local concentration of vacancies in CDAp can be filled with water molecules and thereby raise the c parameter, as observed here in the range of Ca/P < 1.5 (Figure 2b) and in agreement with the STA data (Figure 9c,d). The authors of [44] have demonstrated that CDAp with Ca/P = 1.32–1.48 can be obtained by the precipitation method.
Single-phase samples synthesized by reactions with Ca/P = 1.5–1.6 had apatite structure with varying crystal lattice parameters (Table 2).

4.4. Range 1.67 ≤ Ca/P < 1.90

The formation of single-phase samples with HAp structure is observed in this range (Table 2). The first sample with Ca/P = 1.67 is HAp with a stoichiometric Ca/P value. This sample contains an adsorbed and lattice water (Figure 9f) and a negligible quantity of the carbonate ion absorbed from ambient air during the synthesis (Figure 4d). NMR data show that the signals from phosphate tetrahedra and hydroxyl groups become narrowed, indicating that the environment became more symmetrical. At the same time, in the 1H spectrum, an additional signal appears that can be explained by the exchange between the molecules of adsorbed water and the hydroxyl group of HAp.
With a further increase in Ca/P, the amount of carbonate does not rise (Figure 4d). Nonetheless, lattice parameter a for HAp continues to decrease, while parameter c keeps going up (Figure 2a,b), implying a continued change in the HAp composition. Obviously, due to the high defectiveness of the formed HAp, an increase in Ca/P above 1.67 is possible, and CRIAp arises in this range with the accommodation of excess calcium as an interstitial impurity atom.

4.5. Range Ca/P > 1.90

The samples of this range manifested the presence of calcium hydroxide and oxide, which is a sign of excess calcium in the system. Accordingly, a further increase in calcium concentration in the apatite lattice generated by the mechanochemical synthesis is impossible; however, lattice parameter c of the apatite phase continues to grow in this range (Figure 2b), while a stops diminishing. Consequently, another type of CRIAp is detectable in this range: CRIAp II. The existence of CRIAp is also supported by the data of energy-dispersive X-ray spectroscopy presented in Figure 6c. Stand-alone particles of the sample with Ca/P = 2.0 had a Ca/P ratio close to or equal to the intended one.
NMR signals of phosphate tetrahedra and hydroxyl groups in apatite show a broadening (Table 3 and Table 4), indicating a worsening of the symmetry of the nearest environment. This is probably caused by an increase in the number of hydrogen bonds.
Since CaO particles, which always have a hydroxide shell in the presence of water molecules, are present in this range, we can assume that this entire range is triphasic.

5. Conclusions

Calcium phosphates were synthesized by the mechanochemical method from mixtures of CaHPO4 and CaO with a wide range of initial Ca/P ratios: 1.17 to 2.10. For the first time, the formation of a phase possessing whitlockite structure was documented directly in a ball mill in samples with a low initial Ca/P ratio, 1.17–1.25. This phase contains an HPO42− group and an appreciable amount of structural water. In composition and structure, the identified phase is close to whitlockites synthesized by the hydrothermal method. A hypothesis is advanced that under the conditions of “soft mechanochemical synthesis” in the planetary ball mill, the detected proton-containing whitlockite phase is produced under hydrothermal-like conditions.
A scheme is proposed for several stages of interactions between the components (CaHPO4 and CaO), where at a low Ca/P ratio of 1.17–1.25 and an excess of CaHPO4, whitlockite Ca10HPO4(PO4)6 forms via intermediate products under the hydrothermal conditions. The Ca/P ratio in this compound is 1.428, which can be considered the lowest Ca/P ratio and the highest substitution of calcium by protons, thereby defining the limit of the existence of CDAps transitioning into proton-containing whitlockite.
It was established that in the range 1.33 ≤ Ca/P ≤ 2.10, a continuous series of apatites of varied composition is produced. On the basis of the initial Ca/P ratio, several sets of apatites with characteristic features can be distinguished (Figure 11). CDAp of type I was registered in the range 1.33 ≤ Ca/P ≤ 1.48. This range is biphasic. Upon transition to the single-phase range (1.5 ≤ Ca/P < 1.67), the structure of CDAp alters, CDAp II comes into being, which differs from CDAp I by the absence of structurally bound water. Stoichiometric HAp forms at Ca/P = 1.67. Next comes the range of CRIAp of type I (1.5 < Ca/P ≤ 1.80), which transforms at Ca/P > 1.8 into CRIAp II. The existence of two types of CRIAps is proposed based on the dynamics of changes in their crystal lattice parameters.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ceramics5030031/s1, Video S1: Video S1.avi.

Author Contributions

Conceptualization, M.V.C.; funding acquisition, N.V.B.; investigation, S.V.M., I.Y.P., N.B.K., A.V.I., E.S.P., K.B.G. and O.B.V.; data curation, N.V.B.; methodology, N.V.B. and M.V.C.; writing—original draft preparation, M.V.C.; writing—review and editing, N.V.B. and O.B.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was carried out within a state assignment to the Institute of Solid State Chemistry and Mechanochemistry SB RAS (project No. 121032500064-8).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these results are included in the Materials and Methods section.

Acknowledgments

The TEM analysis was conducted on the equipment of the multi-access center National Center of Catalyst Research (Novosibirsk, Russia).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of the data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Elliott, J.C. Structure and Chemistry of Apatite and Other Calcium Orthophosphates; Studies in Inorganic Chemistry; Elsevier: Amsterdam, The Netherlands, 1994; pp. 1–404. ISBN 0-444-81582-1. [Google Scholar]
  2. Pataquiva-Mateus, A.Y.; Ferraz, M.P.; Monteiro, F.J. Nanoparticles of hydroxyapatite: Preparation, characterization and cellular approach—An Overview. Mutis 2013, 3, 43–57. [Google Scholar] [CrossRef] [Green Version]
  3. Ptáček, P. (Ed.) Synthetic Phase with the Structure of Apatite. In Chemistry: Apatites and their Synthetic Analogues—Synthesis, Structure, Properties and Applications; InTech: Rijeka, Croatia, 2016; pp. 177–244. ISBN 978-953-51-2265-4. [Google Scholar] [CrossRef] [Green Version]
  4. Turon, P.; del Valle, L.J.; Alemán, C.; Puiggalí, J. Biodegradable and Biocompatible Systems Based on Hydroxyapatite Nanoparticles. Appl. Sci. 2017, 7, 60. [Google Scholar] [CrossRef] [Green Version]
  5. Suchanek, W.; Yoshimura, M. Processing and properties of hydroxyapatite-based biomaterials for use as hard tissue replacement implants. J. Mater. Res. 1998, 13, 94–117. [Google Scholar] [CrossRef]
  6. Liu, Q.; Huang, S.; Matinlinna, J.P.; Chen, Z.; Pan, H. Insight into Biological Apatite: Physiochemical Properties and Preparation Approaches. BioMed Res. Int. 2013, 2013, 929748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Combes, C.; Rey, C. Amorphous calcium phosphates: Synthesis, properties and uses in biomaterials. Acta Biomater. 2010, 6, 3362–3378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Dorozhkin, S.V. Calcium orthophospates (CaPO4): Occurrence and properties. Prog. Biomater. 2016, 5, 9–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Mucalo, M. Hydroxyapatite (HAp) for Biomedical Applications; Woodhead Publishing Limited: Waltham, MA, USA, 2015; pp. 1–364. ISBN 978-1-78242-033-0. [Google Scholar] [CrossRef]
  10. Kenny, S.M.; Buggy, M. Bone cements and fillers: A review. J. Mater. Sci. Mater. Med. 2003, 14, 923–938. [Google Scholar] [CrossRef]
  11. Gibson, I.R.; Rehman, I.; Best, S.M.; Bonfield, W. Characterization of the transformation from calcium-deficient apatite to β-tricalcium phosphate. J. Mater. Sci. Mater. Med. 2000, 11, 799–804. [Google Scholar] [CrossRef]
  12. Ghosh, R.; Sarkar, R. Synthesis and characterization of sintered beta-tricalcium phosphate: A comparative study on the effect of preparation route. Mater. Sci. Eng. C 2016, 67, 345–352. [Google Scholar] [CrossRef]
  13. Brown, W.E.; Eidelman, N.; Tomazic, B. Octacalcium phosphate as a precursor in biomineral formation. Adv. Dent. Res. 1987, 1, 306–313. [Google Scholar] [CrossRef]
  14. Tung, M.S.; Brown, W.E. The role of octacalcium phosphate in subcutaneous heterotopic calcification. Calcif. Tissue Int. 1985, 37, 329–331. [Google Scholar] [CrossRef]
  15. Vallet-Regi, M.; Navarrete, D.A. (Eds.) Biological Apatites in Bone and Teeth. In Nanoceramics in Clinical Use: From Materials to Applications, 2nd ed.; Royal Society of Chemistry: London, UK, 2015; pp. 1–29. ISBN 978-1-78262-255-0. [Google Scholar] [CrossRef]
  16. Massit, A.; El Yacoubi, A.; Rezzouk, A.; El Idrissi, B.C. Thermal Behavior of Mg-Doped Calcium-Deficient Apatite and Stabilization of beta Tricalcium Phosphate. Biointerface Res. Appl. Chem. 2020, 10, 6837–6845. [Google Scholar] [CrossRef]
  17. Nakahira, A.; Nakata, K.; Numako, C.; Murata, H.; Matsunaga, K. Synthesis and Evaluation of Calcium-Deficient Hydroxyapatite with SiO2. Mater. Sci. Appl. 2011, 2, 1194–1198. [Google Scholar] [CrossRef] [Green Version]
  18. Andrés-Vergés, M.; Fern, C.; Martínez-Gallego, M.; Solier, J.D.; Cachadiña, I.; Matijević, E. A new route for the synthesis of calcium-deficient hydroxyapatites with low Ca/P ratio: Both spectroscopic and electric characterization. J. Mater. Res. 2000, 15, 2526–2533. [Google Scholar] [CrossRef]
  19. Ergun, C.; Evis, Z.; Webster, T.J.; Sahin, F.C. Synthesis and microstructural characterization of nano-size calcium phosphates with different stoichiometry. Ceram. Int. 2010, 37, 971–977. [Google Scholar] [CrossRef]
  20. Bonel, G.; Heughebaert, J.C.; Heughebaert, M.; Lacout, J.L.; Lebugle, A. Apatite calcium orthophosphates and related compounds for biomaterials preparation. Ann. N. Y. Acad. Sci. 1988, 523, 115–130. [Google Scholar] [CrossRef]
  21. Ansari, M.; Naghib, S.M.; Moztarzadeh, F.; Salati, A. Synthesis and characterization of hydroxyapatite calcium hydroxide for dental composites. Ceramucs-Silikáty 2011, 55, 123–126. [Google Scholar]
  22. Tamai, M.; Nakamura, M.; Isshiki, T.; Nishio, K.; Endoh, H.; Nakahira, A. A metastable phase in thermal decomposition of Ca-deficient hydroxyapatite. J. Mater. Sci. Mater. Med. 2003, 14, 617–622. [Google Scholar] [CrossRef]
  23. Chaikina, M.V.; Bulinaa, N.V.; Vinokurova, O.B.; Prosanov, I.Y.; Dudina, D.V. Interaction of calcium phosphates with calcium oxide or calcium hydroxide during the “soft” mechanochemical synthesis of hydroxyapatite. Ceram. Int. 2019, 45, 16927–16933. [Google Scholar] [CrossRef]
  24. Patent of Russian Federation; №975068; Bull. Inv.: Moscow, Russia, 1982; p. 43.
  25. Chaikina, M.V.; Bulina, N.V.; Vinokurova, O.B.; Prosanov, I.Y. Synthesis of Stoichiometric and Substituted β-Tricalciumphosphate using Mechanochemistry. Chem. Sustain. Dev. 2020, 28, 73–78. [Google Scholar] [CrossRef]
  26. Sakka, S.; Bouaziz, J.; Ben, F. Mechanical Properties of Biomaterials Based on Calcium Phosphates and Bioinert Oxides for Applications in Biomedicine. In Advances in Biomaterials Science and Biomedical Applications; Pignatello, R., Ed.; InTech: Rijeka, Croatia, 2013; pp. 23–50. ISBN 978-953-51-1051-4. [Google Scholar] [CrossRef] [Green Version]
  27. Trandafir, D.L.; Mirestean, C.; Turcu, R.V.F.; Frentiu, B.; Eniu, D.; Simon, S. Structural characterization of nanostructured hydroxyapatite–iron oxide composites. Ceram. Int. 2014, 40, 11071–11078. [Google Scholar] [CrossRef]
  28. Legrand, A.P.; Bresson, B.; Guidoin, R.; Famery, R. Mineralization followup with the use of NMR spectroscopy and others. J. Biomed. Mater. Res. 2002, 63, 390–395. [Google Scholar] [CrossRef]
  29. Boanini, E.; Gazzano, M.; Nervi, C.; Chierotti, M.R.; Rubini, K.; Gobetto, R.; Bigi, A. Strontium and Zinc Substitution in β-Tricalcium Phosphate: An X-ray Diffraction, Solid State NMR and ATR-FTIR Study. J. Funct. Biomater. 2019, 10, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Yashima, M.; Sakai, A.; Kamiyama, T.; Hoshikawa, A. Crystal structure analysis of β-tricalcium phosphate Ca3(PO4)2 by neutron powder diffraction. J. Solid State Chem. 2003, 175, 272–277. [Google Scholar] [CrossRef]
  31. Konishi, T.; Yamashita, K.; Nagata, K.; Lim, P.N.; Thian, E.S.; Aizawa, M. Solid-state nuclear magnetic resonance study of setting mechanism of β-tricalcium phosphate-inositol phosphate composite cements. J. Phys. Mater. 2019, 2, 034007. [Google Scholar] [CrossRef]
  32. Jäger, C.; Welzel, T.; Meyer-Zaika, W.; Epple, M. A solid-state NMR investigation of the structure of nanocrystalline hydroxyapatite. Magn. Reson. Chem. 2006, 44, 573–580. [Google Scholar] [CrossRef] [PubMed]
  33. Mathew, R.; Gunawidjaja, P.N.; Izquierdo-Barba, I.; Jansson, K.; García, A.; Arcos, D.; Vallet-Regí, M.; Edén, M. Solid-State 31 P and 1 H NMR Investigations of Amorphous and Crystalline Calcium Phosphates Grown Biomimetically from a Mesoporous Bioactive Glass. J. Phys. Chem. C. 2011, 115, 20572–20582. [Google Scholar] [CrossRef] [PubMed]
  34. Tõnsuaadu, K.; Gross, K.A.; Plūduma, L.; Veiderma, M. A review on the thermal stability of calcium apatites. J. Therm. Anal. Calorim. 2012, 110, 647–659. [Google Scholar] [CrossRef]
  35. Calvo, C.; Gopal, R. The Crystal Structure of Whitlockite from the Palermo Quarry. Am. Mineral. 1975, 60, 120–133. [Google Scholar]
  36. Britvin, S.N.; Pakhomovskii, Y.A.; Bogdanova, A.N.; Skiba, V.I. Strontiowitlockite, Sr9Mg(PO3OH)(PO4)6, a new mineral species from the Kovdor deposit, Kola Peninsula, USSR. Can. Mineral. 1991, 29, 87–93. [Google Scholar]
  37. Kostiner, E.; Rea, J.R. The crystal structure of manganese-whitlockite, Ca18Mn2H2(PO4)14. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1976, 32, 250–253. [Google Scholar] [CrossRef]
  38. Deyneko, D.V.; Aksenov, S.M.; Morozov, V.A.; Stefanovich, S.Y.; Dimitrova, O.V.; Barishnikova, O.V.; Lazoryak, B.I. A new hydrogen-containing whitlockitetype phosphate Ca9(Fe0.63Mg0.37)H0.37(PO4)7: Hydrothermal synthesis and structure. Z. Kristallographie-Cryst. Mater. 2014, 229, 823–830. [Google Scholar] [CrossRef]
  39. Destainville, A.; Champion, E.; Bernache-Assollant, D.; Laborde, E. Synthesis, characterization and thermal behavior of apatitic tricalcium phosphate. Mater. Chem. Phys. 2003, 80, 269–277. [Google Scholar] [CrossRef]
  40. Bohner, M.; Santoni, B.L.G.; Döbelin, N. β-tricalcium phosphate for bone substitution: Synthesis and properties. Acta Biomater. 2020, 113, 23–41. [Google Scholar] [CrossRef]
  41. Kwon, Y.S.; Gerasimov, K.B.; Yoon, S.K. Ball temperatures during mechanical alloying in planetary mills. J. Alloys Compd. 2002, 346, 276–281. [Google Scholar] [CrossRef]
  42. Boldyrev, V.V. Hydrothermal reactions under mechanochemical action. Powder Technol. 2002, 122, 247–254. [Google Scholar] [CrossRef]
  43. Bulina, N.V.; Chaikina, M.V.; Andreev, A.S.; Lapina, O.B.; Ishchenko, A.V.; Prosanov, I.Y.; Gerasimov, K.B.; Solovyov, L.A. Mechanochemical Synthesis of SiO44−-Substituted Hydroxyapatite, Part II—Reaction Mechanism, Structure, and Substitution Limit. Eur. J. Inorg. Chem. 2014, 2014, 4810–4825. [Google Scholar] [CrossRef]
  44. Christoffersen, J.; Christoffersen, M.R.; Kibalczyc, W.; Andersen, F.A. A contribution to the understanding of the formation of calcium phosphates. J. Cryst. Growth 1989, 94, 767–777. [Google Scholar] [CrossRef]
Figure 1. PXRD patterns of mechanochemically synthesized samples with different Ca/P ratios.
Figure 1. PXRD patterns of mechanochemically synthesized samples with different Ca/P ratios.
Ceramics 05 00031 g001
Figure 2. Dependence of the HAp lattice parameters a (a) and c (b) and crystallite size (c) on the Ca/P ratio.
Figure 2. Dependence of the HAp lattice parameters a (a) and c (b) and crystallite size (c) on the Ca/P ratio.
Ceramics 05 00031 g002
Figure 3. Dependence of the whitlockite lattice parameters a (a) and c (b) on the Ca/P ratio.
Figure 3. Dependence of the whitlockite lattice parameters a (a) and c (b) on the Ca/P ratio.
Ceramics 05 00031 g003
Figure 4. FTIR spectra of the synthesized samples with Ca/P ratios 1.17–1.25 (a), 1.33–1.48 (b), 1.50–1.6 (c), and 1.67–2.10 (d).
Figure 4. FTIR spectra of the synthesized samples with Ca/P ratios 1.17–1.25 (a), 1.33–1.48 (b), 1.50–1.6 (c), and 1.67–2.10 (d).
Ceramics 05 00031 g004aCeramics 05 00031 g004b
Figure 5. TEM (ac) and HRTEM (df) images of samples with Ca/P ratios 1.25 (a,d), 1.42 (b,e), and 2.00 (c,f) at different magnification levels.
Figure 5. TEM (ac) and HRTEM (df) images of samples with Ca/P ratios 1.25 (a,d), 1.42 (b,e), and 2.00 (c,f) at different magnification levels.
Ceramics 05 00031 g005
Figure 6. Distribution maps of elements across a particle in samples with Ca/P ratios 1.25 (a), 1.42 (b), and 2.00 (c).
Figure 6. Distribution maps of elements across a particle in samples with Ca/P ratios 1.25 (a), 1.42 (b), and 2.00 (c).
Ceramics 05 00031 g006
Figure 7. Changes in the HRTEM image of a particle containing bubbles after 1 s (a), 30 s (b), and 60 s (c) of electron beam exposure. Ca/P = 1.42. A video showing these changes is presented as Supplementary Material.
Figure 7. Changes in the HRTEM image of a particle containing bubbles after 1 s (a), 30 s (b), and 60 s (c) of electron beam exposure. Ca/P = 1.42. A video showing these changes is presented as Supplementary Material.
Ceramics 05 00031 g007
Figure 8. 31P (a) and 1H (b) MAS NMR spectra of the synthesized samples with different Ca/P ratios.
Figure 8. 31P (a) and 1H (b) MAS NMR spectra of the synthesized samples with different Ca/P ratios.
Ceramics 05 00031 g008
Figure 9. STA of samples with Ca/P ratios 1.21 (a), 1.25 (b), 1.33 (c), 1.42 (d), 1.50 (e), and 1.67 (f). 1: weight loss; 2: DSC; 3: evolution of water.
Figure 9. STA of samples with Ca/P ratios 1.21 (a), 1.25 (b), 1.33 (c), 1.42 (d), 1.50 (e), and 1.67 (f). 1: weight loss; 2: DSC; 3: evolution of water.
Ceramics 05 00031 g009
Figure 10. FTIR spectra of samples with different Ca/P ratios after STA.
Figure 10. FTIR spectra of samples with different Ca/P ratios after STA.
Ceramics 05 00031 g010
Figure 11. An outline of ranges of existence of the apatites produced by the mechanochemical method of synthesis.
Figure 11. An outline of ranges of existence of the apatites produced by the mechanochemical method of synthesis.
Ceramics 05 00031 g011
Table 1. Reactions of mechanochemical synthesis of calcium phosphates and the intended composition of their products.
Table 1. Reactions of mechanochemical synthesis of calcium phosphates and the intended composition of their products.
Preset
Ca/P Ratio
Composition of Initial MixturePossible Products of Synthesis
1.176CaHPO4 + 1.02CaOTCP, CaHPO4
1.206CaHPO4 + 1.19CaOTCP, CaHPO4
1.216CaHPO4 + 1.23CaOTCP, CaHPO4
1.256CaHPO4 + 1.50CaOTCP, CaHPO4
1.336CaHPO4 + 2CaO + 3H2OOCP
1.386CaHPO4 + 2.28CaOCDAp
1.426CaHPO4 + 2.52CaOCDAp
1.456CaHPO4 + 2.70CaOCDAp
1.486CaHPO4 + 2.88CaOCDAp
1.506CaHPO4 + 3.00CaOCDAp
1.556CaHPO4 + 3.30CaOCDAp
1.606CaHPO4 + 3.60CaOCDAp
1.676CaHPO4 + 4.00CaOOHAp
1.706CaHPO4 + 4.20CaOHAp, Ca(OH)2
1.806CaHPO4 + 4.80CaOHAp, Ca(OH)2
1.906CaHPO4 + 5.40CaOHAp, Ca(OH)2
2.006CaHPO4 + 6.00CaOHAp, Ca(OH)2
2.106CaHPO4 + 6.60CaOHAp, Ca(OH)2
Table 2. Phase composition and structural data of the HAp phase in the synthesized samples.
Table 2. Phase composition and structural data of the HAp phase in the synthesized samples.
Ca/PConcentration (wt.%)Structural Data for HApRwp
CaHPO4CaOCa(OH)2HApβ-TCPAmorphous Apatitea (nm)c (nm)Crystallite Size (nm)
1.173251~170.95040.6911105.2
1.202462~140.95040.6911145.5
1.211963~180.95040.6911156.1
1.251024660.9504(4)0.6911(4)16.1(9)5.9
1.3349510.9474(2)0.6899(2)18.3(8)4.5
1.3849510.9468(2)0.6898(3)18.5(4)5.3
1.4283170.9462(2)0.6893(2)19.3(4)4.9
1.459550.9458(2)0.6892(2)20.0(4)4.9
1.489190.9455(2)0.6891(2)20.6(4)4.7
1.501000.9453(2)0.6889(2)20.6(2)4.6
1.551000.9448(2)0.6889821.3(2)4.6
1.601000.9444(2)0.6891(1)20.6(2)4.5
1.671000.9439(1)0.6893(1)20.4(2)4.4
1.701000.9436(1)0.6894(1)19.6(2)4.3
1.801000.9432(1)0.6897(1)20.6(2)4.4
1.901000.9430(1)0.6898(1)20.9(2)4.3
2.001990.9429(1)0.6900(1)20.1(2)3.5
2.1033940.9430(2)0.6904(2)18.2(2)3.4
Table 3. 31P chemical shift parameters for the synthesized samples.
Table 3. 31P chemical shift parameters for the synthesized samples.
Ca/P PO43− PO43− (HAp) HPO42−Q1−
1.25 δ i s o (ppm)1.20−7.99
Width (Hz)823670
Content0.890.11
1.50 δ i s o (ppm)2.88−7.33
Width (Hz)345884
Content0.960.04
1.67 δ i s o (ppm)4.812.85
Width (Hz)566273
Content0.110.89
2.00 δ i s o (ppm)5.602.98
Width (Hz)466332
Content0.110.89
Table 4. 1H chemical shift parameters for the synthesized samples.
Table 4. 1H chemical shift parameters for the synthesized samples.
Ca/P HPO42−HPO42−H2OH-Bonded OH (HAp) OH (HAp)
1.25 δ i s o (ppm)16.012.266.981.280.08
Width (Hz)130015503400810317
Content0.080.330.540.030.02
1.50 δ i s o (ppm)5.141.35−0.06
Width (Hz)2023170289
Content0.560.010.43
1.67 δ i s o (ppm)5.590.990
Width (Hz)2100710238
Content0.630.100.27
2.00 δ i s o (ppm)4.501.630
Width (Hz)16931153351
Content0.300.320.37
Table 5. Composition of the samples after STA.
Table 5. Composition of the samples after STA.
Ca/PConcentration (wt%)
HApβ-TCPCa2P2O7
1.214753
1.255644
1.337624
1.429010
1.50892
1.67100
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chaikina, M.V.; Bulina, N.V.; Vinokurova, O.B.; Gerasimov, K.B.; Prosanov, I.Y.; Kompankov, N.B.; Lapina, O.B.; Papulovskiy, E.S.; Ishchenko, A.V.; Makarova, S.V. Possibilities of Mechanochemical Synthesis of Apatites with Different Ca/P Ratios. Ceramics 2022, 5, 404-422. https://doi.org/10.3390/ceramics5030031

AMA Style

Chaikina MV, Bulina NV, Vinokurova OB, Gerasimov KB, Prosanov IY, Kompankov NB, Lapina OB, Papulovskiy ES, Ishchenko AV, Makarova SV. Possibilities of Mechanochemical Synthesis of Apatites with Different Ca/P Ratios. Ceramics. 2022; 5(3):404-422. https://doi.org/10.3390/ceramics5030031

Chicago/Turabian Style

Chaikina, Marina V., Natalia V. Bulina, Olga B. Vinokurova, Konstantin B. Gerasimov, Igor Yu. Prosanov, Nikolay B. Kompankov, Olga B. Lapina, Evgeniy S. Papulovskiy, Arcady V. Ishchenko, and Svetlana V. Makarova. 2022. "Possibilities of Mechanochemical Synthesis of Apatites with Different Ca/P Ratios" Ceramics 5, no. 3: 404-422. https://doi.org/10.3390/ceramics5030031

Article Metrics

Back to TopTop