Open access

Biomimetic Modifications of Calcium Orthophosphates

Written By

Diana Rabadjieva, Stefka Tepavitcharova, Kostadinka Sezanova, Rumyana Gergulova, Rositsa Titorenkova, Ognyan Petrov and Elena Dyulgerova

Submitted: 01 November 2010 Published: 29 August 2011

DOI: 10.5772/19054

From the Edited Volume

On Biomimetics

Edited by Assoc. Lilyana D. Pramatarova

Chapter metrics overview

3,925 Chapter Downloads

View Full Metrics

1. Introduction

Calcium orthophosphates are subject to intensive investigations owing to their biological importance. The ion-substituted non-stoichiometric nano-sized poorly crystalline calcium orthophosphates, mainly with apatite structure, build the inorganic component of hard tissues in the organisms. The main ion substitutes are the ions Na+, K+, Mg2+, Fe2+, Zn2+, Si2+, CO3 2-, Cl-, and F- (Dorozhkin, 2009; Daculsi et al., 1997) and they differ in variety and amount depending on the type of the hard tissue, its age as well as on individual peculiarities. The so called “biological apatite” is formed in the living organisms as a result of biomineralization processes, the mechanism of which is not yet clarified. These processes include precipitation, dissolution and growth of poorly-crystalline calcium orthophosphates taking place in the organic matrix, e.g., collagen in the case of bones (Dorozhkin, 2009; Palmer et al., 2008) or amelogenin in the case of enamel (Palmer et al., 2008), in the presence of body fluids. One of the ways to elucidate the elementary processes occurring during bone hard tissue mineralization is the biomimetic approach designed to study these processes. The knowledge of the elementary processes is crucial for the development of new bioactive calcium phosphate materials (close to the natural ones) that may be applied for bone repairing, reconstruction and remodeling.

The aim of this chapter is to throw light on the biomimetic precipitation and modification of calcium orthophosphates, XRD-amorphous calcium phosphate (ACP) and dicalcium phosphate dihydrate (DCPD) on the basis of authors’ kinetic, spectral (XRD and IR) and thermodynamic studies and literature data.

Advertisement

2. Calcium orthophosphates – short review

2.1. Classification

Eleven calcium orthophosphates are known in the literature. According to the methods of their preparation they are divided into two groups - calcium phosphates precipitates and calcium phosphates calcinates (Table 1). The preparation of calcium phosphates precipitates strongly depends on pH of the medium; that of calcium phosphates calcinates is a function of the calcination temperature.

Abbreviation Chemical formula Ca/P Preparation conditions
PRECIPITATES pH
MCPM Ca(H2PO4)2.H2O 0.5 0 - 2
MCPA Ca(H2PO4)2 0.5
DCPD Ca(HPO4)2.2H2O 1.0 2 - 6
DCPA Ca(HPO4)2 1.0
OCP Ca8(PO4)4(HPO4)2.5H2O 1.33 5.5 - 7
ACP Ca9(PO4)6.nH2O 1.5 5 - 12
PCA Ca10-xx(PO4)6-x(HPO4)x.((OH)2-xx) 1.33-1.67 6.5 - 9.5
HA Ca10(PO4)6(OH)2 1.67 9.5 - 12
CALCINATES T,ºC
TCP β-Ca3(PO4)2 1.5 "/800
α-Ca3(PO4)2 1.5 "/1125
ТТCP Ca4(PO4)2О 2.0 "/1500

Table 1.

a the table was adapted according to Dorozhkin (2009), Chow and Eanes (2001) and Johnsson and Nancollas (1992).Calcium orthophosphatesa.

2.2. Structures

Calcium phosphates are divided into three groups according to their structure (Chow & Eanes, 2001): (i) Ca-PO4 sheet-containing compounds (MCPA, MCPM, DCPA DCPD). DCPD has a monoclinic structure, space group Ia, where HPO4 2- ions are linked to Ca2+ ions forming linear chains, that are stacked and form corrugated sheets parallel to the (010) face. The water molecules are situated between the sheets, bonded to the Ca2+ ion. The packing of the Ca-HPO4 ions in chains or sheets determine several possible pseudohexagonal arrangements, similar to the glaserite type structure (Curry & Jones, 1971; Dickens et al., 1972; Dickens & Bowen, 1971); (ii) glaserite type compounds (α-TCP and β-TCP). Two types of columns along the c-axis in a pseudohexagonal arrangement, one containing only Ca2+ and other both Ca2+ and PO4 3- ions in a ratio 1:2 build the glaserite type structure of monoclinic α-TCP (Mathew et al., 1977). In rhombohedral β-TCP structure two types of columns contain both Ca2+ and PO4 3- ions (Dickens et al., 1974). One of the columns has vacancies at both cationic and anionic position; and (iii) apatite type compounds (OCP, TTCP, PCA and HA) (Chow & Eanes 2001; Mathai&Takagi, 2001). Commonly, HA has a hexagonal structure (space group P63/m) (Kay et al., 1964), where Ca2+ ions occupy two different crystallographic symmetry sites. Ca1 are located in columns along the c-axis, where is coordinated to nine O atoms. The Ca-O9 polyhedra are connected in chains parallel to c-axis. Ca2 are arranged in two triangular units. The Ca2 ions are 7-coordinated, with six O atoms and one OH- ion. Ca1 and Ca2 polyhedra are linked through oxygen atoms of the PO4 3- tetrahedra. Each OH- ion occupies statistically disordered positions.

OCP has a triclinic structure, which can be described as alternating along (100) “hydrated” and apatitic layers (Mathew et al., 1988). The atomic positions of the structure of OCP are very close to HA structure, which is the precondition for possible epitaxial growth and formation of interlayered structures, important for explanation of the process of biomineralization.

A special position holds the amorphous calcium phosphate (ACP) which structure is built of Ca9(PO4)6, so called Posner’s clusters, where Ca2+ and PO4 3- ions are arranged in a hexagonal dense packing (Betts et al., 1975; Blumenthal et al., 1977).

The existing symmetry relations between these structures ensure the easier phase transformations.

2.3. Solubility

Calcium orthophosphates are sparingly soluble in water (Table 2). HA has the lowest solubility among them, which is its natural priority. The solubility of calcium phosphates strongly depends on pH of the medium and this feature is of significance for their preparation and biological behavior. Thus, the practically insoluble mono-phase bio-ceramics of dense HA do not actively participate in the process of bone remodeling (Tas, 2004). However, upon contact with body fluids they participate in the formation of a surface layer of bone-like apatite. Mono-phase α-TCP and β-TCP display higher solubilities and rapidly degrade in vitro and in vivo (Radin & Ducheyne, 1993, 1994). Mg- and Zn-doped TCP ceramics display lower solubility than pure TCP ceramics and thus reduce the resorption rate (Xue, 2008). Bi-phase mixtures of HA and β-TCP ceramics were developed in order to improve the biological behaviour of the mono-phase materials (Petrov et al., 2001; Teixeira et al., 2006).

The knowledge on the Ca2+, H+/ OH-, PO4 3-//H2O system and its sub-systems may be used as a theoretical base for predetermination or optimization of the conditions for the preparation of different calcium orthophosphates. Unfortunately, owing to the low solubility and narrow crystallization fields of the different stable and metastable salts, there are no systematic experimental studies of this system. Only single solubility data are available for the binary sub-system Ca2+/PO4 3-//H2O at 25ºC (Kirgintzev et al., 1972). More detailed studies were performed on the three-component Ca2+, H+/ PO4 3-//H2O sub-system and experimental data are available for the temperature range 0 – 100ºC (Flatt et al., 1961; Bassett, 1958; Flatt et al., 1956; Chepelevskii et al., 1955; Belopol’skii, 1940; Bassett, 1917). Two hydrous and two anhydrous salts, namely Ca(H2PO4)2, Ca(H2PO4)2.H2O, CaHPO4 and CaHPO4.2H2O are established at 25ºC and 40ºC respectively; there are contradictions about the existence and stability of the salt of lowest solubility CaHPO4.2H2O (Bassett, 1917; Belopolskii et al., 1940; Chepelevskii et al., 1955). The solubility of Ca(H2PO4)2 and Ca(H2PO4)2.H2O slightly increases at temperatures above 50ºC but CaHPO4.2H2O was not detected (Bassett, 1917; Chepelevskii et al., 1955).

The most appropriate method for evaluation of the solubility of sparingly soluble calcium phosphate salts is the thermodynamic modeling. The ion association model based on the extended Debye-Huckel theory was applied to the Ca2+, H+/ OH-, PO4 3-//H2O system (Chow & Eanes, 2001; Johnnson & Nancollas, 1992). Thermodynamic data for the solubility products (lgKsp 0) of all calcium orthophosphates and the complex formation constants (lgK0) of all complex species which may exist in aqueous calcium phosphate solutions are necessary for its application (Table 2). The calculations of Chow and Eanes (2001) have shown that DCPA is the least soluble salt in the Ca2+, H+/ OH-, PO4 3-//H2O system at pH < 4.2 and 25ºC while HA becomes the least soluble salt at pH > 4.2; TTCP is the most soluble salt at pH <8.2 while DCPD is the most soluble salt at pH > 8.2. In the pH region 7.3 - 7.4 typical for body fluids, the solubility of the salts at 25ºC (Chow & Eanes, 2001) and 37ºC (Johnnson & Nancollas, 1992) follows the order:

TTCP > α-TCP > DCPD > DCPA >OCP ~ β-TCP > HA

Chemical formula Solubility, g/l, 25oC
(Dorozhkin, 2009)
-lgKsp0
Ca(H2PO4)2.H2O ~18 1.14 (Fernandez et al., 1999)
Ca(H2PO4)2 ~17 1.14 (Fernandez et al., 1999)
CaHPO4.2H2O ~0.088 6.59 (Gregory et al., 1970)
CaHPO4 ~0.048 6.90 (McDowell et al., 1971)
Ca8H2(PO4)6.5H2O ~0.0081 96.6 (Tung et al., 1988)
Ca3(PO4)2(am) - 25.2 (Meyer & Eanes 1978)
Ca5(PO4)3ОН ~0.0003 58.4 (McDowell et al., 1977)
α-Ca3(PO4)2 ~0.0025 25.5 (Fowler & Kuroda, 1986)
β-Ca3(PO4)2 ~0.0025 28.9 (Gregory et al., 1974)
Ca4(PO4)2О ~0.0007 38.0 (Matsuya et al., 1996)
Complex formation constants
(National Institute of Standards and Technology [NIST], 2003)
H+ + H2PO4 -=H3PO4 0 2.148
H+ + HPO4 2-=H2PO4 - 7.198
H+ + PO4 3- = HPO4 2- 12.37
Ca2++OH-=CaOH+ 1.303
Ca2++ HPO4 2-=CaHPO4 0 2.66
Ca2++ H2PO4 -=CaH2PO4 + 1.35
Ca2++ PO4 3-= CaPO4 - 6.46

Table 2.

Solubility and thermodynamic data of the Ca2+, H+/ OH-, PO4 3-//H2O system.

Advertisement

3. Electrolyte systems for biomimetic studies

Electrolyte solutions of different composition, designed to mimic the аcellular human body plasma, have become a modern way to test bone-bonding abilities of bioactive materials or to produce thin calcium-phosphate layers on materials (metals, alloys or glasses) for bone graft substitutes (Yang & Ong, 2005; Raghuvir et al., 2006; Jalota et al., 2006; Kontonasaki et al., 2002). The composition of the most popular ones is presented in Table 3.

Earle's balanced salt solution (EBSS, Ca/P = 1.8, HCO3 - - 26.2mmol.dm-3) (Earle et al., 1943) and Hank’s balanced salt solution (HBSS, Ca/P = 1.6, HCO3 - - 4.2mmol.dm-3) (Hanks & Wallace, 1949) were among the first simulated body solutions. Kokubo (1990) was the first to popularize a multicomponent inorganic solution, called conventional simulated body fluid (SBFc) which contains definite amounts of Na+, K+, Mg2+, Ca2+, Cl-, HCO3 2-, HPO4 2- and SO4 2- ions, has a Ca/P ratio of 2.5 (equal to that in the blood plasma), HCO3 - concentration of 4.2 mmol.dm-3 and physiologic pH of 7.3-7.4. To mimic the blood plasma in terms of the most important HCO3 - ions, Bayractar and Tas (1999) revised the SBFc by increasing HCO3 - concentration up to 27 mmol.dm-3 at the account of Cl- ions (revised simulated body fluid, SBFr). The concentrations of Ca2+ and Mg2+ ions in the ionic SBF (SBFi) correspond to those of free Ca2+ and Mg2+ ions (not bound to proteins), in the blood plasma (Oyane, et al., 2003).

Ion content Blood
Plasma
EBSS
(Earle, et al., 1943)
HBSS
(Hanks and Wallace, 1949)
SBFc
(Kokubo, 1990)
SBFr
(Bayraktar and Tas, 1999)
SBFi
(Oyane, et al., 2003)
SBFg
(this study)
Na+ 142.0 143.5 142.1 142.0 142.0 142.0 142.0
K+ 5.0 5.4 5.3 5.0 5.0 5.0 5.0
Ca2+ 2.5 1.8 1.26 2.5 2.5 1.6 2.5
Mg2+ 1.5 0.8 0.9 1.5 1.5 1.0 1.5
Cl- 103.0 123.5 146.8 147.8 125.0 103.0 147.8
HCO3 2- 27.0 26.2 4.2 4.2 27.0 27.0 4.2
HPO4 2- 1.0 1.0 0.78 1.0 1.0 1.0 1.0
Glycine - - - - - - 135.0
SO4 2- 0.5 0.8 0.41 0.5 0.5 1.5 0.5
Ca/P 2.5 1.8 1.62 2.5 2.5 1.6 2.5
pH 7.4 7.2-7.6 6.7–6.9 7.2-7.4 7.4 7.4 7.3

Table 3.

Electrolyte solutions for in vitro experiments, mmol.dm-3.

These solutions were buffered to the pH of blood plasma with TRIS, BITRIS or HEPES buffers.

SBF modified with glycine (SBFg), essential for the biological system amino acid, was prepared on the basis of conventional SBF. Concentration of glycine was thermodynamically calculated so that the contents of free Ca2+ and Mg2+ ions to be analogous to SBFi.

Advertisement

4. Biomimetic precipitation of ion modified precursors

The biomimetic approach which includes precipitation processes of bioactive calcium phosphates in electrolyte medium of simulated body fluids and uses the influence of the medium composition on their formation and phase transformation have attracted extensive research interest (Xiaobo et al., 2009; Hui et al., 2009; Shibli & Jayalekshmi, 2009; Martin et al., 2009), because of their analogy to the biological mineralization processes. In the following, the authors’ studies on the precipitation of ion-modified ACP and DCPD precursors are summarized.

Various crystal chemical and kinetic factors affect the crystallization process. The ion-modified calcium phosphates are mixed crystals (non-stoichiometric compounds), where part of the ions building the crystal unit cell are substituted by other ions. The ability of the admixture ion to adopt the coordination of the substituted ion determines the substitution degree.

To enable ion modification of calcium phosphate precursors with Na+, K+, Mg2+ and Cl- ions we have performed all our studies using conventional SBFc that was modified for each concrete case. Modified calcium-free simulated body fluid (SBFc-Cam) was used as a solvent for K2HPO4 (Solution 1) and phosphorus-free simulated body fluid (SBFc-Pm) was used as a solvent for CaCl2 (Solutions 2 and 5), for CaCl2 and MgCl2 (Solution 3) and for ZnCl2 (Solution 4) (Table 4). In this way preliminary precipitation was avoided. pH of the mixed solutions was adjusted to 7.2-7.4 using 0.1M HCl or 0.05M 2-amino-2-hydroxymethil-1,3-propandiol.

Ion content SBFc-Cam
(Solution 1)
SBFc-Pm
*(Solution 2)
SBFc-Pm
(Solution 3)
SBFc-Pm
(Solution 4)
SBFc-Pm
**(Solution 5)
Na+ 141.9 141.9 141.9 141.9 141.9
K+ 506.4 3.0 3.0 3.0 5.0
Mg2+ 1.5 1.5 1.5 1.5 1.5
Ca2+ - 418.9 - x 418.9 - x 252.1
Me2+ - x x
Cl- 142.8 975.6 -2x 975.6 142.8+2x 642.0
SO4 2- 0.5 0.5 0.5 0.5 0.5
HCO3 - 4.2 4.2 4.2 4.2 4.2
HPO4 2- 251.7 - - - 0.00

Table 4.

*- in the case of ACP precipitation; ** - in the case of DCPD precipitation; 0< x < 83.8 mmol.dm-3.Modified simulated body fluids (SBFs) (mmol.dm-3) used by the authors.

The electrolyte medium provided by SBF plays a crucial role in the precipitation processes and influences the composition of the precipitated product. Precipitation, co-precipitation, ion substitution and ion incorporation reactions simultaneously take place. The cationic and anionic substitutions are mainly responsible for the calcium deficiency of the precipitated ACP precursors. Two methods – fast mixing or continuous co-precipitation of the reagents were applied in these studies. The method of precipitation affected the size, morphology and chemical homogeneity of the precipitate.

SBF-modified XRD-amorphous calcium-deficient phosphate (ACP) (Fig. 1) with a Ca/P ratio of 1.3 or 1.51 (Table 5) due to ion substitution and incorporation of Na+, K+, Mg2+ and Cl- ions from the SBFs at levels close to those of natural enamel, dentin and bone (Dorozhkin, 2009), was precipitated.

Figure 1.

XRD (a) and IR (b) spectra of SBF modified amorphous calcium phosphate.

The fast precipitation was carried out by mixing Solution 1 and Solution 2 (Table 4) at a Ca/P ratio of 1.67 and pH of 11.5 (maintained by 1M KOH) under intense stirring at room temperature. It is known that the fast mixing, the high supersaturation and the presence of Mg2+ and CO3 2- ions provoke the precipitation of an amorphous calcium-deficient product (Sinyaev et al., 2001; Combes & Rey, 2010). The continuous co-precipitation was carried out by mixing Solution 1 and Solution 2 (Table 4) at a rate of 3 ml/min to precipitate in glycine buffer (Sykora, 1976) at room temperature and pH 8 (maintained by 1M KOH).

Mg
mmol/g
Na
mmol/g
K
mmol/g
Cl
mmol/g
Mg/Ca Ca/P (Ca+Mg+Na+K)/P
Biomimetic precipitated ACP at quick mixing
0.13 0.20 0.45 0.03 0.03 1.51 1.79
Biomimetic precipitated ACP at continuous co-precipitation
0.04 0.05 0.01 0.05 0.005 1.3 1.33
Enamel, Dentin, Cementum, Bone (Dorozhkin, 2009)
0.02 - 0.29 0.22 - 0.39 2.10-4 - 0.02 0.03 – 0.1 0.03 – 0.06 1.61 -1.77

Table 5.

Compositions of ACP precursor and natural Enamel, Dentin, Cementum and Bone.

Zn- or Mg-modified amorphous calcium phosphate precursors with varying Me2+/(Ca2++Me2+) ratio from 0.01 to 0.16 (Table 6) due to Ca2+ ion substitution by Me2+ ions as well as Me2+ incorporation were precipitated by the method of continuous co-precipitation in electrolyte system only. All reagents (Solutions 1, 2 and 4 for Zn-modified precursors and Solutions 1 and 3 for Mg-modified precursors, Table 4) with a (Ca2++Me2+)/P ratio of 1.67 (Me2+ = Mg, Zn) were mixed to precipitate in glycine buffer with a rate of 3 ml/min at room temperature and pH 8 (maintained by 1M KOH). The modified conventional simulated body fluids provided ion modification of all Mg- and Zn-modified calcium phosphate precursors with Na+ (0.02 - 0.08 mmol/g), K+ (0.01 – 0.02 mmol/g, Mg 2+ (0.04 mmol/g) and Cl- (below 0.05 mmol/g) ions (Table 6).

Sample Liquid phase Solid phase
Me2+ /(Me 2+ +Ca 2+)
in initial solutions
Me2+ /(Me 2+ +Ca 2+ ) (Ca 2+ +Mg 2+ +Zn 2+ +Na + + K + )/P Zn 2+,
mmol/g
Mg 2+ , mmol/g Na +,
mmol/g
K + , mmol/g Cl-, mmol/g
Zinc-modified calcium phosphates
Zn1 0.01 0.01 1.31 0.09 0.03 0.03 0.01 <0,05
Zn3 0.03 0.03 1.35 0.29 0.05 0.04 0.02 <0,05
Zn5 0.05 0.05 1.35 0.41 0.04 0.05 0.02 <0,05
Zn10 0.10 0.10 1.31 0.90 0.06 0.02 0.01 <0,05
Zn13 0.13 0.13 1.40 1.19 0.05 0.08 0.02 <0,05
Magnesium-modified calcium phosphates
Mg2 0.03 0.02 1.36 - 0.21 0.05 0.02 <0,05
Mg5 0.10 0.05 1.35 - 0.45 0.08 0.01 <0,05
Mg10 0.13 0.10 1.33 - 0.85 0.06 0.02 <0,05
Mg16 0.20 0.16 1.38 - 1.45 0.04 0.02 <0,05

Table 6.

Ion content of the magnesium- and zinc- modified calcium phosphates and their initial solutions.

By analogy with Bigi et al. (1995), we have established that the presence of Zn2+ or Mg2+ ions in the reaction mixture inhibits the crystallization of HA, so that XRD amorphous Mg- or Zn-modified calcium phosphate precursors are obtained. The Posner’s clusters (Betts et al., 1975; Blumenthal et al., 1977) of the complex formula CawMgxZnyNazKu(PO4)v(CO3)6-v (w+x+y+z+u ≤ 9) are the first particles formed in the studied complex electrolyte SBF – CaCl2 – MgCl2/ZnCl2 – KOH – H2O system. A modifying ion, whose ionic radius and electrical charge are closer to those of the Ca2+ ions, will be more readily incorporated into the Pozner’s clusters. The Zn2+ ionic radius (0.74 Å) is closer to that of the Ca2+ ion (1.0 Å) than the radius of the Mg2+ ion (0.65 Å). The substitution with Na+ (0.95 Å) and K+ (1.33 Å) ions is partial not only for geometrical reasons but also for electrostability. The results (Table 6) showed that all Zn2+ ions and only about half of the Mg2+ ions from the reaction solutions were included in the precipitated ACP. The different chemical behavior of Zn2+ and Mg2+ ions can be explained by the “softness-hardness” factor and by the Crystal Field Stabilization Energy (CFSE). According to Pearson’s concept of “hard” and “soft” Lewis acids and bases (Pearson, 1963), as well as the Klopman scale of hardness and softness (Klopman, 1968), “soft acids” predominantly coordinate “soft bases” and “hard acids” - predominantly “hard bases”. Mg is a “hard acid”, while Zn is a “soft acid”. The simulated body fluids contain high concentrations of Cl- ions which are “softer bases” than H2O, OH-, PO4 3-, SO4 2-, HCO3 - and HPO4 2-. Although Zn2+ ions are a “soft acid”, they form a negligible amount of chloride complexes due to the zero value of their CFSE and mainly exist as free Zn2+ ions in the studied solutions. In contrary, Mg2+ as a “hard acid” is preferentially coordinated by the H2O molecules (“hard base”) and are mainly present as [Mg(H2O)6]2+ complexes. The last ones are too large to be incorporated into the crystal structure of the calcium phosphate without its distortion. The necessity of overcoming the energy barrier for even partial dehydration of the [Mg(H2O)6]2+ complexes is another reason for the low substitution rate of these ions.

DCPD biomimetic precipitated precursors - Well crystallized dicalcium phosphate dihydrate (DCPD) (Fig.2) was precipitated by the method of fast mixing (room temperature and intense stirring) of Solution 1 and Solution 5 (Table 4) at a Ca/P ratio of 1 and pH 6 (maintained by 1M HCl). Differently from all modified ACP precursors, only negligible amounts of Mg2+ (0.001 mmol/g), Na+ (0.025 mmol/g), K+ (0.001 mmol/g) and Cl- (0.003 mmol/g) ions were found in biomimetic precipitated DCPD.

Figure 2.

XRD (a) and IR (b) spectra of precipitated DCPD.

Thermodynamic modeling of biomimetic precipitation - The precipitation processes of SBF-modified ACP and DCPD as well as of Zn- and Mg-modified ACP were simulated by an ion-association model using the computer program PHREEQCI v.2.14.3 (Parkhurst, 1995). All possible association/dissociation and dissolution/ crystallization processes in the SBFs were taken into account. The formation of complexes and the precipitation of salts were considered by means of a mass-action expression using the appropriate formation constants or solubility products. The activity coefficients of all possible simple and complex species were calculated by the extended Debye-Huckel theory using an updated database (Todorov, et al., 2006).

The saturation indices (SI) (eq. 1), calculated under the experimental conditions were used as indicators for possible salt crystallization (Table 7),

SI  =  lg ( IAP / K ) E1

where IAP is an ion activity product and K is a solubility product.

When the solution is supersaturated with respect to a certain salt (SI > 0), it will precipitate; when the solution is undersaturated (SI < 0), the salt will not precipitate; the solution and the salt will be in equilibrium when SI = 0.

Different calcium, magnesium, sodium and potassium salts can simultaneously co-precipitate in electrolyte SBF systems. Their number depends on the precipitation conditions (Table 7).

In the SBF with pH value of 11.5, nine salts display positive SI, namely Mg(OH)2, CaHPO4, Mg3(PO4)2.8H2O, MgCO3.Mg(OH)2.3H2O, CaCO3, Ca3(PO4)2(am), Ca8H2(PO4)6.5H2O, Ca9Mg(HPO4)(PO4)6 and Ca10(PO4)6(OH)2 (Table 7) and can co-precipitate. At pH 8 the same salts including CaHPO4.2H2O but except Mg(OH)2 can co-precipitate. The increase of the Mg2+ ion concentration in the system leads to co-precipitation of extra four metastable magnesium salts and favors the precipitation of Ca9Mg(HPO4)(PO4)6 (SI increases). The increase of the Zn2+ concentration in the system does not influence the co-precipitated salts. The only zinc phosphate salt Zn3(PO4)2.4H2O is not expected to precipitate (SI<0). In SBF of pH 6 where DCPD precipitates, only calcium phosphate salts can co-precipitate. In all cases, the highest SI and the highest thermodynamic stability are displayed by Ca10(PO4)6(OH)2 followed by Ca9Mg(HPO4)(PO4)6. Despite the thermodynamic stability of HA, the kinetic factors favor the formation of metastable phases – ACP at pH 8 and 11.5 and initial (Ca+Me)/P = 1.67 and DCPD at pH 6 and Ca/P = 1. These results are in compliance with Ostwald’s step rule, according to which the crystal phase that nucleates is not the phase that is most thermodynamically stable under these conditions, but rather is a metastable phase closest in free energy to the parent phase (Chung, et al., 2009). The highest crystallization rate and the lowest supersaturation necessary for nucleation should be exhibited by those salts in the saturated solution, for which there is a sufficient concentration of structural entities able to be incorporated unchanged or with small changes into the crystal structure.

Advertisement

5. Biomimetic modifications and phase transformations of ACP and DCPD

With the aim to elucidate the influence of micro-environmental surroundings on the phase transformation process of SBF-modified ACP, DCPD, and Zn-modified ACP we have investigated their biomimetic maturation in SBFs by means of kinetic, spectral and thermodynamic studies. The experiments were performed with three different SBFs –

Solid phases SBF modified ACP Mg modified ACP Zn
modified ACP
SBF modified DCPD
pH 11.5 pH 8 pH 8 pH 8 pH 6
NaCl -3.13 -3.18 -3.14 -3.21 -3.29
NaHCO3 -5.75 -3.30 -3.31 -3.24 -3.69
Na2CO3.H2O -6.78 -7.85 -7.82 -7.77 -10.17
Na2CO3.10H2O -4.9 -5.97 -5.94 -5.87 -8.80
NaHCO3.Na2CO3.2H2O -11.65 -10.27 -10.25 -10.13 -12.44
Na2SO4 -6.81 -6.80 -6.82 -6.70 -6.70
Na2SO4.10H2O -5.45 -5.44 -5.46 -5.33 -5.92
KMgPO4:6H2O -0.32 -0.76 0.60 -1.11 -3.32
Mg(OH)2 0.11 -6.63 -5.18 -6.47 -10.95
MgCO3 -0.73 -1.50 -0.11 -1.32 -3.98
MgCO3.3H2O -3.54 -4.31 -2.92 -4.13 -6.49
Mg5(CO3)4(OH)2.4H2O -5.12 -14.94 -7.91 -14.07 -27.17
MgCO3.Mg(OH)2.3H2O 1.09 -6.42 -3.57 -6.08 -12.29
MgSO4.7H2O -6.47 -6.16 -4.81 -5.96 -6.13
MgHPO4.3H2O -3.49 -0.38 0.95 -0.60 -0.76
Mg3(PO4)2 -1.08 -1.60 2.5 -1.90 -6.72
Mg3(PO4)2.8H2O 0.78 0.27 4.36 -0.03 -4.86
Mg3(PO4)2.22H2O -1.23 -1.74 2.35 -2.02 -6.85
Ca(OH)2 -1.39 -8.52 -8.45 -8.54 -12.06
CaCO3 2.62 1.45 1.46 1.46 -0.87
CaSO4 -1.67 -1.76 -1.80 -1.74 -1.67
CaSO4.2H2O -1.44 -1.53 -1.56 -1.50 -1.49
CaHPO4 0.14 2.86 2.86 2.86 2.20
CaHPO4.2H2O -0.15 2.56 2.51 2.17 1.96
Ca3(PO4)2(am) 8.37 6.66 6.62 5.85 1.35
Ca8H2(PO4)6.5H2O 26.63 28.64 28.45 26.24 17.14
Ca9Mg(HPO4)(PO4)6 34.19 32.19 33.39 29.54 15.88
Ca10(PO4)6(OH)2 60.18 47.92 47.87 45.49 28.35
Zn(OH)2 -6.44
ZnCO3 -5.35
ZnCO3.H2O -5.10
Zn2(OH)3Cl -12.28
ZnSO4.H2O -13.98
ZnSO4.6H2O -12.88
ZnSO4.7H2O -12.64
Zn2(OH)2SO4 -16.36
Zn3(PO4)2.4H2O -9.49

Table 7.

Saturation indices (SI) of solid phases in the studied systems.

conventional, SBFc, revised, SBFr, and conventional modified with glycine, SBFg (Table 3). Before maturation the precipitated precursors were filtered, washed with water and with acetone (solid-to-liquid ratio of 1:1) and lyophilized at -56ºC. Then the freeze-dried samples were matured for varying time periods (from 1 h to 6 months) at a solid-to-liquid ratio of 1:250, physiological temperature of 37ºC, in a static regime.

The biomimetic modification of SBF-modified ACP in SBFc, SBFr and SBFg gave rise to changes in the compositions of both solid and liquid phases during the maturation process (Fig. 3) and revealed that dissolution/crystallization processes are strongly influenced by the content of SBFs.

Figure 3.

Kinetic profiles of PO4 3- (a), Ca2+ (b) and Mg2+ (c) in liquid and solid (d, e) phases.

A similar behavior of PO4 3- (Fig. 3a), Ca2+ (Fig. 3b), and Mg2+ (Fig. 3c) ions was found in the three SBFs. The biggest changes were registered during the first 6 hours. The liquid phases were enriched in Ca2+, Mg2+, and PO4 3- ions during the first 2 - 4 h; afterwards, until the 6th hour, the enrichment rate gradually decreased. The highest increase was registered for Ca2+ (32%) and Mg2+ (46%) ions in SBFg. Their highest decrease was observed in SBFr. For Ca2+ ions this decrease was about 90% for 6 hours, while for Mg2+ ions the decrease continued after the 6th hour and reached 83% at the 72nd hour. The presence of glycine in SBFg and the higher content of HCO3 - ions in SBFr leads to formation of metal-glycine and metal-carbonate complexes that enhance the solubility of the salts. During maturation in SBFr which is richer in HCO3 -, crystallization of CaCO3 occurs, also confirmed by the increased Ca/P ratios in the solid phase (Fig.3d), whereas in SBFc and SBFg the formation of calcium phosphate dominates. The increase of the Mg/Ca ratio in the solid phases (Fig.3e) gives an evidence for the incorporation of Mg2+ in the amorphous phase.

The spectral studies (XRD and IR) confirmed the biomimetic phase transformation of amorphous calcium phosphate into the more stable poorly-crystalline apatite in the three SBFs, differing only in the phase transformation rate (Fig 4). Crystal phase was detected at the 4th hour of the maturation process in SBFc; at the 2nd hour in SBFr and at the 1st hour in SBFg (Fig 4a). The increase in the degree of crystallinity during the maturation process was confirmed by the observed splitting of the phosphate bands at 960, 1100, 562, and 603 cm-1, which are characteristic for the IR spectra of crystalline calcium phosphate (Fig 4b).

Figure 4.

XRD patterns (a) and IR (b) spectra for different SBFs and maturation times.

IR data also revealed a change in the carbonate content of the samples treated in SBFc and SBFr with different carbonate content (Fig. 5). As a measure of the carbonate content we used the ratio between the areas underneath the peaks corresponding to CO3 2- (1549-1336 cm-1) and PO4 3- (1280-914 cm-1) stretching bands. As can be seen, the amount of carbonate ions faster increases in samples matured in SBFr than in those matured in SBFc. Thermodynamic modeling of the maturation process of two different solid calcium phosphate products - a metastable amorphous product (*ACP), and a stable equilibrium product (**ACP) were done in the three solutions (SBFc, SBFr and SBFg) differing in their ionic content (Table 8).

Phase Maturation of *ACP Maturation of **ACP
SBFc SBFr SBFcg SBFc SBFr SBFcg
NaCl -3.59 -3.66 -3.59 -3.59 -3.65 -3.59
Na2SO4 -6.15 -6.14 -6.16 -6.14 -6.13 -6.14
Na2SO4.10H2O -5.34 -5.32 -5.35 -5.33 -5.31 -5.33
NaHCO3 -3.53 -3.17 -3.59 -3.81 -2.4 -3.54
NaHCO3.Na2CO3.2H2O -9.76 -8.69 -9.94 -12.95 -9.05 -12.04
Na2CO3.H2O -7.66 -6.94 -7.78 -10.56 -8.08 -9.93
Na2CO3.10H2O -6.25 -5.53 -6.37 -9.15 -6.67 -8.53
MgSO4.7H2O -5.26 -5.84 -5.33 -5.25 -5.25 -5.22
MgCO3 -1.17 -1.05 -1.34 -4.07 -1.6 -3.40
MgCO3.3H2O -3.67 -3.55 -3.84 -6.56 -4.1 -5.90
Mg3(PO4)2 -2.27 -1.90 -2.78 -9.29 -5.77 -8.00
MgHPO4.3H2O -1.01 -0.89 -1.19 -1.91 -1.21 -1.64
Mg3(PO4)2.22H2O -2.3 -1.93 -2.83 -9.32 -5.8 -8.06
Mg5(CO3)4(OH)2.4H2O -10.95 -10.33 -11.80 -27.75 -15.76 -24.36
KMgPO4.6H2O -2.92 -2.43 -3.25 -6.43 -4.66 -5.91
Ca(OH)2 -9.01 -9.01 -9.01 -14.34 -13.49 -14.06
CaSO4 -2.77 -3.48 -2.66 -2.86 -4.13 -3.29
CaSO4.2H2O -2.58 -3.28 -2.48 -2.67 -3.94 -3.10
CaMg3(CO3)4 -3.18 -2.81 -3.69 -14.86 -6.25 -12.67
CaHPO4.2H2O -0.23 -0.23 -0.23 -1.23 -1.8 -1.42
Mg(OH)2 -5.95 -5.93 -6.16 -11.05 -9.0 -10.47
Mg3(PO4)2.8H2O -0.37 0 -0.89 -7.14 -3.77 -6.16
MgCO3.Mg(OH)2.3H2O -4.41 -4.27 -4.83 -12.29 -7.75 -11.12
CaCO3 0 0 0 -2.94 -1.77 -2.80
CaHPO4 0 0 0 -1.05 -1.61 -1.11
Ca3(PO4)2(am) 0 -0.10 0 -6.76 -7.64 -7.43
Ca8H2(PO4)6.5H2O -6.62 -8.43 -7.19
Ca9Mg(HPO4)(PO4)6 -11.37 -11.55 -11.36
Ca10(PO4)6(OH)2 0 0 0

Table 8.

Notes: The impurities in the washed initial products were taken in the range 1-3 % based on the measured Mg/Ca ratio (3 mol %) (Table 5).*ACP – metastable amorphous product; **ACP – stable equilibrium product.Likely salts in the biomimetic systems and their thermodynamic calculated saturated indices (SI) at biomimetic maturation.

Figure 5.

Changes in the carbonate content of the samples treated in SBFc and SBFr.

The metastable amorphous product (*ACP) simulated the system behavior during the first 1-2 hours of maturation, when no Ca8H2(PO4)6.5H2O, Ca9Mg(HPO4)(PO4)6 and Ca10(PO4)6(OH)2 phase was yet formed, whereas the stable product (**ACP) simulated the equilibrium system. The maturation of the metastable amorphous product (*ACP) leads to a phase transformation that depends on the content of HCO3 - ions in SBF at the beginning of the process (Table 8). In a solution with a low HCO3 - content (SBFc and SBFg), dissolution phenomena of all magnesium salts occur (SI<0) during maturation and the system will be in equilibrium with the calcium salts (SI = 0), including the amorphous calcium phosphate. The increase in the HCO3 - content (SBFr) leads to dissolution and phase transformation of the amorphous calcium phosphate into thermodynamically more stable salts. The calculated equilibrium amounts of CaCO3 and Ca3(PO4)2(am) in the three investigated body fluids (Fig. 6) point to the significantly favored crystallization of CaCO3 (especially in SBFr) and the decreased amount of Ca3(PO4)2(am) due to dissolution processes. The calculations revealed that there was no influence of SBFs composition on the equilibrium product (**ACP), the system tending to thermodynamic equilibrium by dissolution of all co-precipitated solid phases and re-crystallization of the thermodynamically unstable amorphous calcium phosphate (with SI<0) into pure HA (with SI = 0) (Table 8).

Figure 6.

Calculated equilibrium amounts of CaCO3 and Ca2(PO4)3(am).

The model cannot predict the formation of Mg-substituted carbonated hydroxyapatite due to the lack of thermodynamic data.

The calculated species distribution in the initial SBFs, in the SBFs at metastable (maturation of *ACP) and at stable (maturation of **ACP) equilibrium (Fig. 7) gives an evidence for the domination of Me2+ (Me = Ca, Mg) free ions in all studied cases. In the initial SBFs (Fg. 7a and 7b) and at a stable thermodynamic equilibrium (Fig. 7e and 7f) Me2+ free ions are dominating, followed by MeCl+ in SBFc, from MeHCO3 - and MeCl+ in SBFr and from CaH(Gly)2+ and CaCl+ and from MgCl+ and Mg(Gly)+ in SBFg. Significant changes in species distribution are observed at a metastable equilibrium (Fig. 7c and 7d), revealing the essential role of SBF ionic composition on the maturation process. The increased amount of MeHPO4 2-, CaPO4 - in SBFc and SBFr, as well as the increased amount of Me(Gly)+ species in SBFg is due to the dissolution of metastable salts while the decreased amount of MeHCO3 - species especially in SBFr is due to the crystallization of CaCO3.

These thermodynamic data explain the results from the maturation kinetics.

The biomimetic modifications of Zn-modified ACP were studied on three exemplary samples with different Zn2+/(Zn2+ + Ca2+) molar ratios (0.03, 0.05 and 0.10), treated in a conventional simulated body fluid (SBFc). It was found that the Zn content decreases by a factor of 2 during the first 2 hours (Fig. 8) when the samples are still amorphous (Fig. 9). Subsequently, the amorphous phase progressively converted into poorly-crystalline apatite. The Zn content influenced the transformation rate. At a higher Zn content the stability of the amorphous phase increased and the rate of the process slowed down (Fig. 9).

The kinetic studies of the biomimetic modifications of DCPD revealed that the compositions of the liquid and solid phases, similar to those of ACP, changed during the maturation process (Fig. 10). In SBFg the highest increase in PO4 3- and Ca2+ concentrations was registered as a result of the effect of glycine which promotes the dissolution of DCPD. In the solid phase the Ca/PO4 ratio is kept ≈1 during the first 6 hours, then gradually increases. The latter is an indication for the beginning of the transformation process and the formation of basic calcium phosphates with Ca/PO4> 1.

Figure 7.

Ca and Mg species distribution in initial SBF (a and b); at maturation of *ACP (c and d); at maturation of **ACP (e and f)

Figure 8.

Time dependence of the solid phase Zn2+/(Zn2+ + Ca2+) molar ratio during maturation.

Figure 9.

XRD patterns of solid phases with Zn2+/(Zn2+ + Ca2+) ratios of 0.03 (a); 0.05 (b) and 0.1 (c) matured for different times in SBFc.

Figure 10.

Kinetic profiles of pH (a), PO4 3- (b), Ca2+ (c), in the liquid phases and of the Ca/PO4 ratio in the solid phases (d) during maturation.

IR and XRD data (Fig.11) show that the initial DCPD transforms into poorly-crystalline B-type carbonate apatite via an intermediate phase of OCP. The rate of conversion of DCPD to carbonated apatite differs, depending on the type of SBF. The formation of carbonated apatite is faster in SBFr, as it is indicated by the disappearance of O-H and P-OH peaks typical of DCPD phase and the appearance of absorption bands near 1420 and 1480 cm-1 associated with B-type CO3 2- groups after 240 h of treatment (Rabadjieva et al., 2010). For samples treated in SBFc DCPD phase is still preserved after 72 h of treatment, while the carbonate peaks appeared after 720h (one month) maturation and increased in intensity after 6 months, indicating slow rate of PO4 3− substitution by CO3 2- groups. Intense absorption bands in the 1073-1122 cm-1 spectral range are due to the asymmetric stretching vibration mode of P-O, whereas the absorption band that appears near 963 cm-1 is associated with the symmetric stretching mode of P-O. Peaks at 602, 562 and at 470 cm-1 originate from O-P-O bending modes. O-H stretching bands of crystalline water appeared in the range of 3500-3100 cm-1, while H2O bending is at 1640-1650 cm-1. Peaks at 1297 and 1192 cm-1 are due to the P-OH bending mode of HPO4 2- groups, while those near 914 and 870 cm-1 arise from the stretching mode of HPO4 2- groups and partially overlapped with C-O vibrations. The absorption bands at around 878 cm-1 could be due to bending vibration of CO3 2- groups, because are related with intensive bands in the 1400-1500 cm-1 spectral interval, typical of C-O stretching vibrations. Peak at 526 cm-1 is associated with HO–PO3 bending mode in HPO4 2- (Mandel & Tas, 2010; LeGeros et al., 1989).

Figure 11.

IR spectra (a) and XRD patterns (b) at different maturation time.

Advertisement

6. High-temperature phase transformations of ion modified ACP

The matured Zn- and Mg-modified ACP precursors were treated by a procedure including gelation with Xanthan gum, lyophilization at -56ºC, washing (solid-to-water ratio of 1:100) and secondary lyophilization. Then they were sintered at 600, 800 and 1000ºС in order to study the high-temperature phase transformations and to follow the effect of Mg2+ and Zn2+ on the Ca2+ ion substitution and on the phase composition and morphology of the sintered products. The working regime consisted of heating at a rate of 3ºC/min up to the desired temperature and keeping it constant for 1 hour. Both, concentration of Mg2+ and Zn2+ ions and temperature affect the spectral characteristics of the studied samples (Fig. 12). The hydroxyl stretching peaks near 3570 cm-1 and the hydroxyl librations at 633 cm-1 revealed the presence of HA phase, while the peaks near 1120, 1075, 1040, 975 and 950 cm-1 indicated the formation of β-TCP phase.

Figure 12.

FT IR spectra of Zn-modified (a) and Mg-modified (b) calcium phosphates at different temperatures.

The XRD data (Fig. 13) revealed that all amorphous precursors were transformed into ion-modified HA or β-TCP crystalline phases depending on the additive and its amount as well as on the temperature. Both Zn and Mg substitutions promoted the ACP transformation to ion-modified HA and β-TCP but the effect was more pronounced in the case of Zn substitution. Zn-β-TCP and Mg-β-TCP were registered at 600ºC for samples with Me2+/(Me2+ + Ca2+) ratios higher than 0.05 and mixture of the ion-modified HA and β-TCP calcium phosphates at ratios lower than 0.05. The SEM analyses revealed that Zn and Mg substitutions influenced the morphology of the β-TCP grains (Fig 14). Sintering at 800-1000ºC leads to zinc-modified-β-TCP powders of idiomorphic crystals with sizes ranging from 500 to 5000 nm in all studied cases. Magnesium-modified-β-TCP fine powders with spherical grains of smaller size (250 - 1000 nm) were obtained at a Mg2+/(Mg2+ + Ca2+) ratio higher than 0.02 (Figure 13 and 14). The surface of the spherical grains of magnesium modified samples was covered by blanket, while clean grain boundary was observed in the zinc-modified samples.

Figure 13.

XRD powder data of Zn-modified (a) and Mg-modified (b) calcium phosphates at different temperatures. (▄ - HAP; not marked – Me2+-β-TCP).

The calculated unit cell parameters showed that low ion substitution (Me2+/(Me2++Ca2+) ratio up to 0.05) leads to a slight decrease in the parameter a and the volume V and to an increase in the parameter c of the Me-β-TCP unit cell (Table 9). This tendency became more pronounced at higher temperatures.

The inclusion of Mg2+ and Zn2+ ions into the crystal unit cell of the thermodynamically stable HA proceeds through Ca2+ ion substitution. As the ionic radii of Mg2+ (0.65 Å) and

Sample 600ºC 800ºC 1000ºC
a, Å c, Å V, Å3 a, Å c, Å V, Å3 a, Å c, Å V, Å3
Zinc-modified calcium phosphates
Zn1 * * * 10.423 37.339 3513 10.417 37.393 3514
Zn3 * * * 10.437 37.101 3500 10.390 37.142 3471
Zn5 * * * 10.397 37.062 3469 10.387 37.042 3460
Zn10 10.345 36.902 3420 10.329 36.903 3410 10.349 37.153 3446
Zn13 10.332 37.005 3421 10.349 36.988 3431 10.339 37.227 3446
Magnesium-modified calcium phosphates
Mg2 * * * 10.420 37.278 3505 10.415 37.301 3504
Mg5 10.419 37.314 3508 10.403 37.324 3498 10.393 37.319 3491
Mg10 10.356 37.161 3452 10.368 37.234 3466 10.346 37.134 3442
Mg16 10.332 37.152 3435 10.340 37.213 3446 10.391 37.28 3486

Table 9.

*poorly crystallized phase.Unit cell parameters of Me2+-modified calcium phosphates.

Zn2+ (0.74 Å) are too small in comparison with that of Ca2+ (1.00 Å) the increase in their amount leads to unit cell distortion and volume decrease, established also by Ito et al. (2002) and Miyaji et al. (2005). Thus, the structure of Me2+ ion modified HA is destabilized and

Figure 14.

SEM images of Me2+-modified calcium phosphates heated at 1000ºC: a) Mg2+/(Mg2++Ca2+) = 0.02; b) Mg2+/(Mg2++Ca2+) = 0.10; c) Zn2+/(Zn2++Ca2+) = 0.01; d) Zn2+/(Zn2++Ca2+) = 0.13.

its transformation into the more stable β-TCP structure could be expected. The latter structure includes different CaOn coordination polyhedra (n=3,6,7,8) (Yashima et al., 2003). The vacant sites of the smallest CaO3 polyhedron are the most suitable holes for the inclusion of the small Mg2+ and Zn2+ ions, thus the unit cell distortion and structure destabilization will be negligible. Ion substitution at a Me2+/(Ca2++Me2+) ratio higher than 0.05-0.15 leads to an increase in the Me2+ ion inclusion into the larger CaOn polyhedra (n=6-8), which destabilizes the structure. The appearance of a more stable high-temperature modification, α-TCP, could be expected in this case, but no α-TCP XRD peaks were detected in our experiments.

Advertisement

7. Conclusions

In a summary, original authors’ studies and literature data are presented on the biomimetic synthesis of XRD-amorphous calcium phosphate and dicalcium phosphate dihydrate and their biomimetic modifications and phase transformations into poorly-crystalline apatite in three types of simulated body fluids - conventional (SBFc), revised (SBFr) and modified with glycine (SBFg). The compositions of the different types of artificial body fluids that are known in the literature are compared in terms of their similarity to blood plasma; their advantages and disadvantages are highlighted. The authors’ studies and original results on chemical and phase compositions, kinetics and thermodynamic simulations are discussed. A new approach based on thermodynamic modeling (using the PHREEQCI v.2.14.3 computer program based on an ion-association model) was applied for simulation and explanation of the biomimetic precipitation of metastable XRD-amorphous calcium phosphate and dicalcium phosphate dihydrate instead of the thermodynamically stable hydroxyapatite and of their biomimetic phase transformations during the maturation processes. The crucial role of the SBF as an electrolyte system is emphasized.

Advertisement

Acknowledgments

This work is financially supported by the Bulgarian Ministry of Education, Youth and Science under Projects DTK 02-70/2009 and CVP-09-0003.

References

  1. 1. Bassett H. 1917The phosphates of calcium. Part IV. The basic phosphates. J. Chem. Soc. Trans., 111 620 642 DOI:CT9171100620
  2. 2. Bassett H. 1958The Phosphates of Calcium, Part V. Revision of the Earlier Space Diagram. J. Chem. Soc., 2949 2955 DOI:JR9580002949
  3. 3. Bayraktar D. Tas A. C. 1999Chemical preparation of carbonated calcium hydroxyapatite powders at 37°C in urea-containing synthetic body fluids. J Eur Ceram Soc, 19 2573 2579 0955-2219
  4. 4. Belopol’skii A. P. Serebrennikova M. T. Belevich A. V. 1940Vapour Pressure of Saturated Solutions and Solubilities in the System CaO-P2O5-H2O (In Russian). Zhurnal Prikl. Khim., 13 1 3 10C.A. 34:77181
  5. 5. Betts F. Blumenthal N. C. Posner A. S. Becker G. L. Lehninger A. L. 1975Atomic Structure of Intracellular Amorphous Calcium Phosphate Deposits. Proc Natl Acad Sci USA, 72 2088 2092 1091-6490
  6. 6. Bigi A. Foresti E. Gandolfi M. Gazzano M. Roveri N. 1995Inhibiting Effect of Zinc on Hydroxylapatite Crystallization, Journal of Inorganic Biochemistry. 58 49 58 0162-0134
  7. 7. Blumenthal N. C. Betts F. Posne,r A. S. 1977Stabilization of Amorphous Calcium Phosphate by Mg and ATP. Calcif Tissue Res., 23 245 250 0008-0594
  8. 8. Chepelevskii M. L. Bol’ts Ts. S. Vasilenko N. A. Rubinova S. S. 1955Solubility and Rate of Dehydratation of CaHPO4.2H2O (In Russian). Issledovaniya po Priklad. Khim., Akad. Nauk S.S.S.R., Otdel Khim. Nauk.,. 175 183C.A. 50:46156b
  9. 9. Chow L. C. Eanes E. D. 2001 Octacalcium Phosphate, Karger, 380557228Basel, Switzerland
  10. 10. Chung S. Y. Kim Y. M. Kim J. G. Kim Y. J. 2009Multiphase transformation and Ostwald’s rule of stages during crystallization of a metal phosphate. Nature Physics, 5 68 73 1745-2473
  11. 11. Combes C. Rey C. 2010Amorphous calcium phosphates: Synthesis, properties and uses in biomaterials, Acta Biomaterialia, 6 3362 3378 1742-7061
  12. 12. Curry N. A. Jones D. W. 1971Crystal Structure of Brushite, Calcium Hydrogen Orthophosphate Dihydrate: A Neutron-Difraction Investigation. J Chem. Soc., (A), 3725 3729 DOIJ19710003725
  13. 13. Daculsi G. Bouler J. M. Le Geros R. Z. 1997Adaptive crystal formation in normal and pathological calcifications in synthetic calcium phosphate and related biomaterials. Int. Rev. Cytology, 172 129 191 0074-7696
  14. 14. Dickens B. Bowen J. S. 1971Refinement of the Crystal Structure of Ca(H2PO4)2.H2O. Acta Crystallogr., B27 2247 2255ISNN 1600-5740
  15. 15. Dickens B. Bowen J. S. Brown W. E. 1972A Refinement of the Crystal Structure of CAHPO4 (Synthetic Monetite). Acta Crystallogr., B28 797 806ISNN 1600-5740
  16. 16. Dickens B. Schroeder L. W. Brown W. E. 1974Crystallographic Studies of the Role of Mg as a Stabilizing Impurity in b-Ca3(PO4)2. I. The cryastal structure of pure b-Ca3(PO4)2. J. Solid State Chem, 10 232 248ISNN 0022-4596
  17. 17. Dorozhkin S. V. . April 2009Calcium Orthophosphates in Nature, Biology and Medicine. Materials, 2April 2009), 399 498 1996-1944
  18. 18. Earle W. R. Schilling E. L. Stark T. H. Straus N. P. Brown M. F. Shelton E. 1943Production of malignancy in vitro. IV. The mouse fibroblast cultures and changes seen in the living cells. JNCI, 4 165 212 0027-8874
  19. 19. Fernandez E. Gil F. J. Ginebra M. P. Driessens F. C. M. Planell J. A. Best S. M. 1999Calcium phosphate bone cements for clinical applications Part I: Solution chemistry. J Mater Sci: Mater Med, 10 169 176ISNN 0957-4530.
  20. 20. Flatt R. Brunisholz G. Dagon R. 1961Contribution a l’etude du systeme quinaire Ca++- NH4+- H+- 3-PO4H2O XXI. Les solutions saturees a 0 et a 50o du systeme quaternaire Ca++- NH4+- H+- PO4---- H2O. Helvetica Chimica Acta, Vol. XLIV, No VII, 2173 2193ISNN 1522-2675
  21. 21. Flatt R. Brunisholz G. Denereaz A. 1956Contribution a l’etude du systeme quinaire Ca++- NH4+- H+- 3-PO4H2O XVIII. Le systeme quaternaire Ca++- H+- NO3-- PO4---- H2O. Helvetica Chimica Acta, Vol. XXXIX, No II, 473 483ISNN 1522-2675
  22. 22. Fowler B. O. Kuroda S. 1986Changes in Heated and in Laser-Irradiated Human Tooth Enamel and their Probable Effects of Solubility. Calcif Tissue Int., 38 197 208 0017-1967X
  23. 23. Gregory T. M. Moreno E. C. Petel J. M. Brown W. E. 1974Solubility of b-Ca3(PO4)2 in the System Ca(OH)2- H3PO4 at 5, 15, 25 and 37oC. J. Res. Nat. Bur. Stand., 78A 667 674ISNN 0022-4332
  24. 24. Gregory T. M. Moreno E. C. Brown W. E. 1970Solubility of CaHPO4.2H2O in the System Ca(OH)2-H3PO4-H2O at 5, 15, 25, 37.5oC. J Res Natl Bur Stand, 74A 461 475ISNN 0022-4332
  25. 25. Hanks J. H. Wallace R. E. 1949Relation of oxygen and temperature in the preservation of tissues by refrigeration. Proc. Soc. Exp. Biol. Med., 71 196 200 1525-1373
  26. 26. Hui W. Changjian L. Ren H. 2009Effects of structure and composition of the CaP composite coatings on apatite formation and bioactivity in simulated body fluid. Applied Surface Science, 255 4074 4081 0169-4332
  27. 27. Ito A. Kawamura H. Otsuka M. Ikeuchi M. Ohgushi H. Ishikawa K. Onuma K. Kanzaki N. Sogo Y. Ichinose N. 2002Zinc-releasing calcium phosphate for stimulating bone formation. Materials Science and Engineering, C22 21 25ISNN 0921-5093
  28. 28. Jalota S. Bhaduri S. B. Tas A. C. 2006Effect of carbonate content and buffer type on calcium phosphate formation in SBF solutions. J Mater Sci:Mater Med., 17 697 707ISNN 0957-4530.
  29. 29. Johnsson-A M. S. Nancollas G. H. 1992The role of Brushite and Octacalcium Phosphate in Apatite Formation. Critical Reviews in Oral Biology and Medicine, 3No ½, 61 82ISNN 1045- 4411
  30. 30. Kay M. I. Young R. A. Posner A. S. 1964Crystal Structure oh Hydroxyapatite. Nature, 204 1050 1052ISNN 0028-0836
  31. 31. Kirgintzev A. N. Trushnikova L. N. Lavrent’eva V. G. 1972 Solubility of Inorganic Substances in Water (in Russian), Khimia, UDK 541.8(083), Leningrad, Russia
  32. 32. Klopman G. 1968Chemical reactivity and the concept of charge- and frontier-controlled reactions. J. Am. Chem. Soc., 90 2January, 1968), 223 234ISNN 0002- 7863.
  33. 33. Kokubo T. 1990Surface Chemistry of Bioactive Glass-Ceramics. J Non-Cryst Solids, 120 138 151 0022-3093
  34. 34. Kontonasaki E. Zorba T. Papadopoulou L. Pavlidou E. Chatziztavrou X. Paraskevopoulos K. Koidis P. 2002Hydroxy carbonate apatite formation on particulate bioglass in vitro as a function of time, Cryst Res Technol, 37 1165 1171ISNN 0232-1300
  35. 35. Le Geros R. Z. Daculsi G. Orly I. Abergas T. Torres W. 1989Solution-mediated transformation of octacalcium phosphate (OCP) to apatite. Scanning Microsc.,3 1March, 1989), 129 138 0891-7035
  36. 36. Mandel S. Tas A. C. 2010Brushite (CaHPO4 2H2O) to octacalcium phosphate (Ca8(HPO4)2(PO4)4 5H2O) transformation in DMEM solutions at 36.5 °C. Materials Science and Engineering, Vol. C 30, 245 254ISNN 0921-5093
  37. 37. Martin R. A. Twyman H. Qiu D. Knowles J. C. Newport R. J. 2009A study of the formation of amorphous calcium phosphate and hydroxyapatite on melt quenched Bioglass using surface sensitive shallow angle X-ray diffraction. J Mater Sci: Mater Med, 20 883 888ISNN 0957-4530.
  38. 38. Mathew M. Takagi S. 2001Structures of Biological Minerals in Dental Research. J. Res. Natl. Inst. Stand. Technol., 106 6November-December 2001), 1035 1044 0160-1741
  39. 39. Mathew M. Brown W. E. Schroeder L. W. Dickens B. 1988Crystal Structure of Octacalcium Bis(Hydrogenphosphate)Tetrakis(Phosphate) Pentahydrate, Ca8(HPO4)2(PO4)4.5H2O. J. Chem. Crystallogr., 18 235 250 1074-1552
  40. 40. Mathew M. Schroeder L. W. Dickens B. Brown W. E. 1977The crystal Structure of a-Ca3(PO4)2. Acta Crystallogr., B33 1325 1333ISNN 1600-5740.
  41. 41. Matsuya S. Takagi S. Chow L. C. 1996Hydrolysis of tetracalcium phosphate in H3PO4 and KH2PO4. J. Mater. Sci., 31 3263 3269ISNN 0022-2461
  42. 42. Mc Dowell H. Brown W. E. Sutter J. R. 1971Solubility Study Of Calcium Hydrogen Phosphate. Ion-Pair Formation. Inorganic Chemistry. 10 8 1638 1643ISNN 0020-1669.
  43. 43. Mc Dowell H. Gregory T. M. Brown W. E. 1977Solubility of Ca5(PO4)3 in the System Ca(OH)2- H3PO4- H2O at 5, 15, 25 and 37oC. J. Res. Nat. Bur. Stand., 81A 273 281ISNN 0022-4332
  44. 44. Meyer J. L. Eanes D. E. 1978A Thermodynamic Analysis of the Secondary Transition in the Spontaneous Precipitation of Calcium Phosphate. Calcif. Tissue Res., 25 209 216 0008-0594
  45. 45. Miyaji F. Kono Y. Suyama Y. 2005Formation and structure of zinc-substituted calcium hydroxyapatite. Materials Research Bulletin, 40 209 220ISNN 0025-5408
  46. 46. NIST Standard Reference Database 46, version 7 2003 NIST Critically Selected Stability Constants of Metal Compounds, MD 20899 Gaithersburg, USA.
  47. 47. Oyane A. Kim H. M. Furuya T. Kokubo T. Miyazaki T. Nakamura T. . May 2003Preparation and assessment of revised simulated body fluids. J.Biomed. Mater. Res., 65A 188 195 1552-4965
  48. 48. Palmer L. C. Newcomb C. J. Kaltz S. R. Spoerke E. D. Stupp S. I. 2008Biomimetic Systems for Hydroxyapatite Mineralization Inspired By Bone and Enamel. Chem. Rev., 108 11 4754 4783PMID 19006400
  49. 49. Parkhurst D. L. 1995User’s guide to PHREEQC-A computer program for speciation, reaction-path, advective-transport, and inverse geochemical calculations. U.S. Geological Survey Water-Resources Investigations Report 95 4227 http://wwwbrr.cr.usgs.gov/projects/GWC_coupled/phreeqci/
  50. 50. Pearson R. 1963Hard and soft acids and bases, J. Am. Chem. Soc., 85 22November, 1963), 3533 3539ISNN 0002- 7863
  51. 51. Petrov O. E. Dyulgerova E. Petrov L. Popova R. 2001Characterization of calcium phosphate phases obtained during the preparation of sintered biphase Ca-P ceramics. Materials Letters, 48 162 167ISNN 0167-577X
  52. 52. Rabadjieva D. Gergulova R. Titorenkova R. Tepavitcharova S. Dyulgerova E. Balarew Chr. Petrov O. 2010Biomimetic transformations of amorphous calcium phosphate: kinetic and thermodynamic studies. J Mater Sci: Mater Med., 21 9 2501 2509 1573-4838
  53. 53. Radin S. R. Ducheyne P. 1993Effect of bioactive ceramic composition and structure on in vitro behavior. II. Precipitation. J Biomed Mater Res, 27 35 44ISNN 1549-3296.
  54. 54. Radin S. R. Ducheyne P. 1994Effect of bioactive ceramic composition and structure on in vitro behavior. III. Porous versus dense ceramics. J Biomed Mater Res., 28 1303 1309ISNN 1549-3296.
  55. 55. Raghuvir S. Kurella A. Narendra B. D. 2006Laser Surface Modification of Ti-6Al-4V: Wear and Corrosion Characterization in Simulated Biofluid. J Biomater App, 21 49 73ISNN 0885-3282
  56. 56. Shibli S. M. Jayalekshmi A. C. 2009A novel nano hydroxyapatite-incorporated Ni-P coating as an effective inter layer for biological applications. J Mater Sci: Mater Med, 20 711 718ISNN 0957-4530
  57. 57. Sinyaev V. A. Shustikov,a E. S. Levchenko L. V. Sedunov A. A. 2001Synthesis and dehydration of amorphous calcium phosphate. Inorganic Materials, 37 619 622 0020-1685
  58. 58. Sykora V. 1976 Chemicko analyticke tabulky, DT 543(083), Praha, Czech Republic.
  59. 59. Tas A. C. 2004Participatioon of Calcium Phosphate Bone Substitutesin the Bone Remodeling Process Influence of Materials Chemistry and Porosity. Key Engineering Materials, 264-268 1969 1972ISNN 1662-9795
  60. 60. Teixeira C. C. Nemelivsky Y. Karkia C. Legeros R. Z. 2006Biphasic calcium phosphate: a scaffold for growth plate chondrocyte maturation. Tissue Eng., 12 8August, 2006), 2283 2289
  61. 61. Todorov T. Rabadjieva D. Tepavitcharova S. 2006New thermodynamic database for more precise simulation of metal species in natural waters. Journal of the University of Chemical Technology and Metallurgy. 41 97 102ISNN 1311-7629
  62. 62. Tung M. S. Eidelman N. Sieck B. Brown W. E. 1988Octacalcium phosphate solubility product from 4 to 37oC. J Res Natl Bur Stand, 93 613 624ISNN 0022-4332
  63. 63. Xiaobo Ch. Yuncang L. Peter D. H. Cui’e W. 2009Microstructures and bond strengths of the calcium phosphate coatings formed on titanium from different simulated body fluids. Materials Science and Engineering, Vol. C 29, 165 171 0928-4931
  64. 64. Xue W. Dahlquist K. Banerjee A. Bandyopadhyay A. Bose S. 2008Synthesis and characterization of tricalcium phosphate with Zn and Mg based dopants. J Mater Sci: Mater Med, 19 2669 2677ISNN 0957-4530.

Written By

Diana Rabadjieva, Stefka Tepavitcharova, Kostadinka Sezanova, Rumyana Gergulova, Rositsa Titorenkova, Ognyan Petrov and Elena Dyulgerova

Submitted: 01 November 2010 Published: 29 August 2011