Energetics of Sn2+ isomorphic substitution into hydroxylapatite: first-principles predictions

Philippe F. Weck*a and Eunja Kimb
aSandia National Laboratories, Albuquerque, NM 87185, USA. E-mail: pfweck@sandia.gov
bDepartment of Physics and Astronomy, University of Nevada Las Vegas, Las Vegas, NV 89154, USA

Received 5th September 2016 , Accepted 3rd November 2016

First published on 4th November 2016


Abstract

The energetics of Sn2+ substitution into the Ca2+ sublattice of hydroxylapatite (HA), Ca10(PO4)6(OH)2, has been investigated within the framework of density functional theory. Calculations reveal that Sn2+ incorporation via coupled substitutions at Ca(II) sites is energetically favourable up to a composition of Sn6Ca4(PO4)6(OH)2, and further substitutions at Ca(I) sites proceed once full occupancy of Ca(II) sites by Sn2+ is achieved. Compositions of SnxCa10−x(PO4)6(OH)2 (x = 4–9) are predominant, with an optimal stoichiometry of Sn8Ca2(PO4)6(OH)2, and Sn-substituted HA follows approximately Vegard's law across the entire composition range.


1 Introduction

Hydroxylapatite [Ca10(PO4)6(OH)2], fluorapatite [Ca10(PO4)6(F)2], and chlorapatite [Ca10(PO4)6(Cl)2] end-members of the apatite group are among the most extensively studied minerals of the apatite supergroup, with generic crystal-chemical formula IXM(I)VII4M(II)6(IVTO4)6X2, where M and T represent species-forming cations (M = Ca2+, Pb2+, Ba2+, Sr2+, Mn2+, Na+, Ce3+, La3+, Y3+, Bi3+; T = P5+, As5+, V5+, Si4+, S6+, B3+), X denotes column monovalent anions (X = OH, F, Cl), and left superscripts indicate ideal coordination numbers.1,2 Archetypal apatite minerals belong to the hexagonal-dipyramidal class and crystallize in the P63/m space group – although lower monoclinic symmetry (space group P21/b) can result from orientational ordering of X anions within the [00z] apatic channels.3,4 In P63/m-structured apatites, IXM(I) sites (Wyckoff positions 4f) generally accommodate smaller cation, while VIIM(II) sites (Wyckoff positions 6h) are occupied by larger cations.

In particular, hydroxylapatite (HA) is found in igneous and metamorphic environments and in biogenic deposits,5,6 and is notorious for its facile ion-exchange properties. Owing to its chemical similarity to inorganic bone matrix material, excellent biocompatibility, and favourable surface properties, HA is one of the major components – along with β-tricalcium phosphate, β-Ca3(PO4)2 – of biphasic calcium phosphate (BCP) bioactive ceramics used for bone grafting, implant and prosthesis coatings, and reconstructive surgery.6–10

The mechanisms and energetics of ion uptake and retention in pristine and doped HA have been subjects of active research in recent years, due to the crucial roles in mediating biological processes played by trace ions found in bones (e.g. Sr2+, Zn2+, Mn2+, F, SiO44− and CO32−),6,11–16 and to potential HA and doped-HA applications in catalysis17,18 and immobilization of toxic heavy metals and radionuclides in contaminated soil and ground water.12,19–23

Compared to other metal cations such as Sr2+ or Zn2+ which are widely used in to promote bone formation or retard bone absorption, limited information is available on the energetics of Sn2+ incorporation into HA. An early study by Klement and Haselbeck24 reported the synthesis of Sn8Ca2(PO4)6Cl2 and Sn5Ca5(PO4)6F2 apatites, followed by the preparation of Sn9Ca(PO4)6(OH,Cl,F)2 apatite by McConnell and Foreman.25 Insertion of Sn2+ in dentin (consisting of ∼45 wt% HA) was also recently investigated as a pre-treatment for preventing dental erosive wear, since there is empirical evidence that Sn2+ increases bond strength.26 In addition, Sn(II)-doped HA has been proposed as a potential material for removal and immobilization of 99Tc contaminant (t1/2 = 2.13 × 105 years, β = 294 keV; produced from the nuclear fuel cycle with 6.1 and 5.9% fission yields from the fission of 235U and 239Pu) in wastewater or nuclear waste streams/tanks.23,27,28 Indeed, early experimental ion-exchange studies demonstrated the role of Sn(II) as a reductant of soluble pertechnetate anion (TcVIIO4) in the presence of phosphate or diphosphonate ligands.29,30 Despite the moderate sorption or retention of 99/99mTc in pristine HA,22,31 tin(II)-doped HA was shown to efficiently reduce Tc(VII) and subsequently immobilize Tc(IV), without further reoxidation to volatile Tc2O7.23,27,28

In this work, the energetics of Sn2+ substitution for Ca2+ in HA has been investigated within the framework of density functional theory (DFT), by predicting the evolution of the excess energy of Sn-doped HA as a function of the composition. This study aims at predicting the impact of Sn(II) substitutions in HA on the structure of bioactive ceramics and candidate materials based on Sn-doped HA for removal and immobilization of radionuclides. Combined Sn2+ substitutions at Ca(I) and Ca(II) sites have been systematically assessed and the most energetically favourable compositions of SnxCa10−x(PO4)6(OH)2 (x = 1–10) have been determined. Details of the computational methods utilized in this study are given in the next section, followed by a complete analysis and discussion of our results. A summary of our findings and conclusions is presented in the last section of the manuscript.

2 Computational methods

Total-energy calculations were carried out at 0 K using DFT, as implemented in the Vienna ab initio simulation package (VASP).32 The exchange–correlation energy was calculated using the generalized gradient approximation (GGA), with the parameterization of Perdew, Burke and Ernzerhof (PBE).33 Standard functionals, such as the PBE or PW91 functionals, were found in previous studies to accurately describe the structural parameters and properties of metal–oxygen systems characterized experimentally.34–40 Previous first-principles approaches, using for example GTO/B3LYP, have also been successful in describing HA.41,42

The projector augmented wave (PAW) method was utilized to describe the interaction between valence electrons and ionic cores.43,44 In the Kohn–Sham (KS) equations, the Ca(3p6,4s2), P(3s2,3p3), O(2s2,2p4), and Sn(5s2,5p2) electrons were treated explicitly as valence electrons and the remaining core electrons together with the nuclei were represented by PAW pseudopotentials. The KS equation was solved using the blocked Davidson45 iterative matrix diagonalization scheme, followed by the residual vector minimization method. The plane-wave cutoff energy for the electronic wavefunctions was set to 500 eV, ensuring the total energy of the system to be converged to within 1 meV per atom. Electronic relaxation was performed with the conjugate gradient method accelerated using the Methfessel–Paxton Fermi-level smearing46 with a Gaussian width of 0.1 eV.

The hexagonal structure of HA (space group P63/m, IT no. 176; Z = 2) refined by Hughes et al.47 using single-crystal X-ray diffraction (XRD) was used as an initial guess in structural optimization calculations. Ionic relaxation calculations of SnxCa10−x(PO4)6(OH)2 (x = 0–10) periodic unit cells were carried out using the quasi-Newton algorithm and Hellmann–Feynman forces acting on atoms were calculated with a convergence tolerance set to 0.01 eV Å−1. Structural optimization was performed without symmetry constraints. The Brillouin zone was sampled using the Monkhorst–Pack k-point scheme48 with a 3 × 3 × 3 k-point mesh for all simulated structures.

3 Results and discussion

3.1 Crystal structure of HA

The crystal unit cell of HA optimized with DFT at the GGA/PBE level of theory is shown in Fig. 1 and the corresponding unit-cell parameters are summarized in Table 1, along with results from crystallographic XRD measurements and previous DFT calculations. The calculated cell dimensions of a = b = 9.542 Å, c = 6.897 Å (c/a = 0.723, α = β = 90.0°, γ = 120.0°) are in good agreement with the values of a = b = 9.432 Å, c = 6.881 Å (c/a = 0.729, α = β = 90.0°, γ = 120.0°) measured by Deer et al.,49 and also consistent with the values of Ellis and co-workers19 (a = b = 9.560 Å, c = 6.863 Å; c/a = 0.718, α = β = 90.0°, γ = 120.0°) and de Leeuw50 (a = b = 9.563 Å, c = 6.832 Å; c/a = 0.714, α = β = 90.0°, γ = 120.0°) computed with DFT/GGA.
image file: c6ra22249h-f1.tif
Fig. 1 Crystal unit cell of hydroxylapatite, Ca10(PO4)6(OH)2 (space group P63, IT no. 173; Z = 2), relaxed with DFT at the GGA/PBE level of theory. Top: View along the [001] direction of the apatic channels (c-axis); bottom: side view of the unit cell. Color legend: Ca(I), purple; Ca(II), light blue; P, green; O, red; H, white. Ca(Ia) and Ca(Ib) are labelled distinctively.
Table 1 Calculated and measured structural unit-cell parameters of bulk hydroxylapatite, Ca10(PO4)6(OH)2
  a (Å) c (Å) α (°) γ (°)
a Ellis et al., 2006; ref. 19.b deLeeuw, 2001; ref. 50.c Deer et al., 1992; ref. 49.d Hughes et al., 1989; ref. 47.e McConnell et al., 1966; ref. 25.
This work 9.542 6.897 90 120
DFT(GGA)a 9.560 6.863 90 120
DFT(GGA)b 9.563 6.832 90 120
Expt.c 9.432 6.881 90 120
Expt.d 9.4166 6.8745 90 120
Expt.e 9.416(2) 6.883(2) 90 120


As discussed in details by deLeeuw,50 OH groups are preferentially ordered into ⋯O–H⋯O–H⋯ chains along the [001] direction of the apatic channels (c-axis). As a result, the mirror-plane symmetry through the Ca(II) triangle plane of the HA unit cell is broken and the space group symmetry is effectively reduced from P63/m (IT no. 176) to P63 (IT no. 173). While all six Ca(II) forming apatic channels remain equivalent in the P63 space group, the four Ca(I) split into pairs of Ca(Ia) and Ca(Ib) (cf. Fig. 1). The computed [measured]47 Ca(II)–O distances are 2.35, 2.36, 2.37, 2.41, 2.48, 2.52 and 2.82 Å [2.343(×2), 2.353, 2.385, 2.509(×2) and 2.710 Å]. The calculated distances between Ca(Ia) and its nine first, second and third oxygen atom neighbours are 2.44(×3), 2.47(×3) and 2.80(×3) Å and the predicted Ca(Ib)–O distances are 2.41(×3), 2.44(×3) and 2.88(×3) Å, i.e., in agreement with the measured distances of 2.403(×3), 2.452(×3) and 2.802(×3) Å.47 The calculated P–O bond lengths of 1.55(×3) and 1.56 Å are in line with the experimental estimates of 1.530(×2), 1.534, 1.537 Å.47

3.2 Sn substitution into HA

Typical ion-exchange in HA involves, first, adsorption of divalent metal cations onto the HA surface, followed by cationic substitution into the Ca2+ sublattice of HA.51,52 The reaction corresponding to Sn2+/Ca2+ substitution into HA can be expressed as:
 
Ca10(PO4)6(OH)2 + xSn2+ → Ca10−xSnx(PO4)6(OH)2 + xCa2+, (1)
where x = 1–10.

In the following, the endpoints of pristine hydroxylapatite and fully-substituted Sn-hydroxylapatite are referred to as CaHA and SnHA, respectively, and the normalized composition of Sn-substituted hydroxylapatite is simply written as SnxCa1−xHA (x = 0.1–1). In addition, square-bracket notations [p + q] indicate occupancies p and q on adjacent Ca(II) triangles, as shown in Fig. 2 depicting Ca(II)/Sn(II) cationic arrangements in apatic channels for normalized compositions of Sn-substituted hydroxylapatite, SnxCa1−xHA (x = 0.2–0.4). Sn2+ substitutions at Ca(Ia) and Ca(Ib) sites are designated as Sn(Ia) and Sn(Ib), respectively.


image file: c6ra22249h-f2.tif
Fig. 2 Ca(II)/Sn(II) cationic arrangements in apatic channels for normalized compositions of Sn-substituted hydroxylapatite, SnxCa1−xHA (x = 0.2–0.4). Square-bracket notations [p + q] indicate occupancies p and q on adjacent Ca(II) triangles. Color legend: Ca, light blue; Sn, light purple.

In order to determine the most energetically favourable compositions of the solid solution SnxCa1−xHA (x = 0.1–0.9) relative to the CaHA and SnHA reference end-members, the excess energy was calculated as:

 
Ex = E(SnxCa1−xHA) − xE(SnHA) − (1 − x)E(CaHA), (2)
where negative values of Ex correspond to energetically favourable configurations of the SnxCa1−xHA solid solution. The resulting Ex calculated for various configurations of one to nine Sn2+/Ca2+ substitutions at Ca(I) and Ca(II) sites within the HA unit cell are displayed in Fig. 3, with their corresponding numerical values reported in Table 2. As shown in Fig. 3, substitutions occur preferentially at Ca(II) sites up to a composition of Sn0.6Ca0.4HA and further substitutions at Ca(I) sites proceed only once full occupancy of Ca(II) sites by Sn2+ is achieved (solid black line). This finding is similar to previous Rietveld analyses of XRD data on Pb2+- or (Pb2+,Sr2+)-substituted HA synthesized in aqueous environment.19,53,54 Let us note that the lowest excess-energy configurations [1 + 0] and [2 + 1]-mer for Sn0.1Ca0.9HA and Sn0.3Ca0.7HA, respectively, are not energetically favourable (Ex > 0), while Sn2+ incorporation via coupled substitutions at Ca(II) sites for the [1 + 1]-trans configuration of Sn0.2Ca0.8HA is essential energy neutral. Compositions of SnxCa1−xHA (x = 0.4–0.9) are predicted to be predominant (Ex < 0), with an optimal stoichiometry of Sn0.8Ca0.2HA, and optimal configurations of [3 + 1] (x = 0.4), [3 + 2] (x = 0.5), [3 + 3] (x = 0.6), [3 + 3] + 1 Sn(Ia) (x = 0.7), [3 + 3] + 1 Sn(Ia) + 1 Sn(Ib) (trans) (x = 0.8) and [3 + 3] + 2 Sn(Ia) + 1 Sn(Ib) (x = 0.9).


image file: c6ra22249h-f3.tif
Fig. 3 Excess energy, Ex, as a function of the normalized composition of Sn-substituted hydroxylapatite SnxCa1−xHA (x = 0.1–1) unit-cell calculated at the GGA/PBE level of theory. Legend: combined Sn substitutions at Ca(I) and Ca(II) sites, green triangles; Sn substitution at Ca(I) sites only, blue diamonds; Sn substitution at Ca(II) sites only, red circles; pristine Ca-hydroxylapatite (CaHA), pink square; fully-substituted Sn-hydroxylapatite (SnHA), black square. The most energetically favourable configurations at fixed compositions are connected by a solid black line. Square-bracket notations [p + q] correspond to neighbour configurations, as shown in Fig. 2.
Table 2 Excess energy (Ex) and lattice parameters of SnxCa1−xHA (x = 0–1) unit-cell structures optimized at the GGA/PBE level of theory. Square-bracket notations [p + q] indicate Sn occupancies p and q on adjacent Ca(II) triangles and Sn substitutions at Ca(Ia) and Ca(Ib) sites are referred to as Sn(Ia) and Sn(Ib), respectively
Composition Sn(I) Sn(II) Configuration Ex (eV) a (Å) b (Å) c (Å) V3)
0% 0 0 CaHA 0.0 9.542 9.542 6.897 543.9
10% 0 1 [1 + 0] 0.108 9.619 9.637 6.916 553.9
1 0 [0 + 0], 1 Sn(Ia) 0.244 9.571 9.571 6.947 551.2
1 0 [0 + 0], 1 Sn(Ib) 0.277 9.567 9.567 6.960 551.7
20% 0 2 [1 + 1]-trans 0.003 9.714 9.771 6.896 563.5
0 2 [1 + 1]-cis 0.132 9.729 9.694 6.922 563.3
0 2 [2 + 0] 0.217 9.667 9.668 6.949 562.1
1 1 [1 + 0], 1 Sn(Ia) 0.308 9.650 9.657 6.967 560.8
1 1 [1 + 0], 1 Sn(Ib) 0.361 9.638 9.657 6.964 560.3
2 0 2 Sn(Ia) 0.378 9.576 9.576 7.021 557.5
2 0 1 Sn(Ia), 1 Sn(Ib) (trans) 0.415 9.587 9.587 7.015 558.3
2 0 2 Sn(Ib) 0.477 9.592 9.592 7.005 558.1
2 0 1 Sn(Ia), 1 Sn(Ib) (cis) 0.538 9.600 9.600 6.993 558.2
30% 0 3 [2 + 1]-mer 0.061 9.704 9.755 6.953 571.9
0 3 [2 + 1]-fac 0.155 9.749 9.710 6.984 571.8
1 2 [1 + 1]-trans, 1 Sn(Ia) 0.246 9.764 9.826 6.940 572.6
0 3 [3 + 0] 0.345 9.690 9.701 7.008 569.9
1 2 [1 + 1]-cis, 1 Sn(Ia) 0.367 9.758 9.714 6.979 570.9
1 2 [1 + 1]-cis, 1 Sn(Ib) 0.410 9.760 9.721 6.976 571.2
40% 0 4 [3 + 1] −0.189 9.883 9.740 7.022 583.4
0 4 [2 + 2]-cis 0.032 9.819 9.785 7.009 582.6
0 4 [2 + 2]-trans 0.082 9.797 9.783 7.002 579.9
1 3 [2 + 1]-mer, 1 Sn(Ia) 0.279 9.757 9.790 7.002 581.6
1 3 [2 + 1]-mer, 1 Sn(Ib) 0.284 9.746 9.794 6.998 581.7
2 2 [1 + 1]-trans, 1 Sn(Ia), 1 Sn(Ib) (trans) 0.313 9.800 9.791 7.063 580.6
1 3 [2 + 1]-fac, 1 Sn(Ia) 0.333 9.799 9.749 7.039 580.7
1 3 [2 + 1]-fac, 1 Sn(Ib) 0.387 9.794 9.745 7.032 580.4
2 2 [1 + 1]-trans, 2 Sn(Ia) 0.400 9.780 9.750 7.058 578.1
50% 0 5 [3 + 2] −0.320 9.772 9.794 7.110 588.5
1 4 [2 + 2]-trans, 1 Sn(Ia) 0.120 9.811 9.797 7.106 592.1
1 4 [2 + 2]-cis, 1 Sn(Ia) 0.157 9.861 9.810 7.078 592.7
60% 0 6 [3 + 3] −0.603 9.741 9.740 7.231 594.2
1 5 [3 + 2], 1 Sn(Ia) −0.334 9.762 9.857 7.239 604.5
1 5 [3 + 2], 1 Sn(Ib) −0.282 9.789 9.808 7.209 600.1
2 4 [2 + 2]-trans, 1 Sn(Ia), 1 Sn(Ib) (trans) 0.028 9.825 9.712 7.230 603.7
2 4 [2 + 2]-trans, 2 Sn(Ia) 0.071 9.799 9.753 7.222 604.6
2 4 [2 + 2]-cis, 1 Sn(Ia), 1 Sn(Ib) (trans) 0.154 9.868 9.775 7.199 604.2
2 4 [2 + 2]-cis, 2 Sn(Ia) 0.242 9.842 9.784 7.189 602.4
70% 1 6 [3 + 3], 1 Sn(Ia) −0.600 9.769 9.769 7.336 606.4
1 6 [3 + 3], 1 Sn(Ib) −0.540 9.767 9.765 7.339 606.4
80% 2 6 [3 + 3], 1 Sn(Ia), 1 Sn(Ib) (trans) −0.695 9.773 9.775 7.512 621.5
2 6 [3 + 3], 2 Sn(Ia) −0.663 9.774 9.777 7.498 620.6
2 6 [3 + 3], 1 Sn(Ia), 1 Sn(Ib) (cis) −0.343 9.810 9.809 7.403 617.0
90% 3 6 [3 + 3], 2 Sn(Ia), 1 Sn(Ib) −0.610 9.839 9.850 7.620 639.6
3 6 [3 + 3], 1 Sn(Ia), 2 Sn(Ib) −0.350 9.816 9.815 7.552 630.1
100% 4 6 SnHA 0.0 9.867 9.868 7.646 644.8


These computational results are overall consistent with the reported synthesis of Sn8Ca2(PO4)6Cl2 and Sn5Ca5(PO4)6F2 apatites by Klement and Haselbeck,24 as well as with the preparation of Sn9Ca(PO4)6(OH,Cl,F)2 apatite by McConnell and Foreman.25 Although the fully-substituted SnHA end-member was predicted to crystallize in the P63 space group, isomorphic to CaHA, the formation of SnHA is comparatively less likely to occur owing to its higher energy cost.

The computed cell dimensions of Sn9Ca(PO4)6(OH)2 are a = 9.84 Å, b = 9.85 Å, c = 7.62 Å, i.e. larger than the XRD estimates of a = b = 9.45 Å, c = 6.89 Å determined by McConnell and Foreman25 for Sn9Ca(PO4)6(OH,Cl,F)2 apatite. Interestingly, as found in the latter experimental study, calculations predict a volume expansion as a result of Sn/Ca substitutions in HA, although McConnell and Foreman pointed out that a simple rigid-ion model suggests instead a slight reduction of the unit cell size, based on the comparative Ahrens ionic radii (A-IR) of Sn2+ (0.93 Å) and Ca2+ (0.99 Å).55 However, several caveats to the aforementioned A-IR rigid-ion model prediction are worth considering. Indeed, no coordination number (CN) or polyhedral distortion were accounted for in this A-IR model, where the radii are independent of the structure type. A subsequent CN-dependent rigid-ion model by Shannon and Prewitt56 introduced effective ionic radii (‘IR’) [and crystal radii (CR)], with values of ‘IR’(Ca2+) = 1.00–1.35 Å [CR(Ca2+) = 1.14–1.49 Å] for CN = 6–12 and ‘IR’(Sn2+) = 1.22 Å [CR(Sn2+) = 1.36 Å] for CN = 8. A revision of this ‘IR’ model including polyhedral distortion provided refined estimates of ‘IR’(Ca2+) = 1.00–1.34 Å [CR(Ca2+) = 1.14–1.48 Å] for CN = 6–12 and ‘IR’(Sn2+) = 1.22–1.32 Å [CR(Sn2+) = 1.36–1.46 Å] for CN = 7–9.57

An additional complication arises from the poorly-defined coordination environment of Ca2+ in HA, as discussed in previous studies.58,59 In fact, the coordination environment of Ca(II) can be described as either 4 (∼2.36 Å) + 2 (∼2.51 Å) + 1 (∼2.71 Å) or 5 (∼2.35 Å) + 2 (∼2.51 Å) + 1 (∼2.71 Å), while the coordination environment of Ca(I) is best described as 6 (∼2.43 Å) + 3 (∼2.79 Å).59 Owing to the 2[thin space (1/6-em)]:[thin space (1/6-em)]3 ratio of Ca(I)[thin space (1/6-em)]:[thin space (1/6-em)]Ca(II) in HA, a simplified, single CN of 8.4 was proposed, assuming a relaxed 2.9 Å cutoff for the Ca–O first-shell coordination – a shorter cutoff of 2.6 Å would result in CN = 6. Since ‘IR’(Ca2+) = 1.00–1.18 Å [CR(Ca2+) = 1.14–1.32 Å] for CN = 6–9, and ‘IR’(Sn2+) = 1.22–1.32 Å [CR(Sn2+) = 1.36–1.46 Å] for CN = 7–9, a volume increase of SnxCa1−xHA unit cells concomitant with Sn substitutions for Ca(I)/Ca(II) in HA is therefore predicted by the ‘IR’ model.

The evolution of the optimized unit-cell volume of SnxCa1−xHA (x = 0–1) as a function of the composition is shown in Fig. 4, for all the configurations listed in Table 2. Calculations predict a nearly linear increase of the unit-cell volume of the most energetically favourable configurations. SnxCa1−xHA (x = 0.1–0.6) equilibrium structures resulting from substitutions at Ca(II) sites closely follow Vegard's law for ideal mixture up to x = 0.4, followed by a breakaway in volume increase for x = 0.5 ([3 + 2] configuration) and x = 0.6 ([3 + 3] configuration) from Vegard's predictions. Subsequent Sn(Ia) and Sn(Ib) substitutions tend to gradually bring the SnxCa1−xHA solid solution back to an ideal Vegard's mixing behaviour. It is also worth noting that among the compositions of SnxCa1−xHA (x = 0.4–0.9) predicted to be predominant (Ex < 0), the volume corresponding to the optimal stoichiometry of Sn0.8Ca0.2HA ([3 + 3] + 1 Sn(Ia) + 1 Sn(Ib)) (trans) configuration is the closest to Vegard's prediction within this compositional range. In addition, at low Sn content (see x = 0.1–0.2 in Fig. 4), the most energetically unfavourable structures with Sn substitutions at Ca(I) sites only (cf. Table 2 and Fig. 3) exhibit the largest departure from their corresponding Vegard's volumes. These findings suggest a strong interplay between structure and energetics in SnxCa1−xHA solid solutions across the complete compositional range.


image file: c6ra22249h-f4.tif
Fig. 4 Evolution of the unit-cell volume of SnxCa1−xHA (x = 0–1) as a function of the normalized composition calculated at the GGA/PBE level of theory. Legend: combined Sn substitutions at Ca(I) and Ca(II) sites, green triangles; Sn substitution at Ca(I) sites only, blue diamonds; Sn substitution at Ca(II) sites only, red circles; pristine Ca-hydroxylapatite (CaHA), pink square; fully-substituted Sn-hydroxylapatite (SnHA), black square. The most energetically favourable configurations at fixed compositions are connected by a solid black line. The dashed black line represents the empirical Vegard's law for an ideal mixture.

The density of states for the most energetically favourable structures from pure HA to fully-substituted Sn-HA were also computed in order to understand the electronic structure underlying the physicochemical properties of these compounds. As shown in Fig. 5, all the structures are semiconductors, with a band gap decreasing from 5.3 eV for pure HA to 3.0 eV for fully-substituted Sn-HA. The top of the valence space is dominated by O(2p) states, with increasing O(2p)–Sn(5s) hybridization with Sn doping. The bottom of the conduction space possesses predominantly Sn(5p) character. Both Ca and P electronic states are not expected to have significant impact on the physicochemical properties of these compounds since their electronic states are much deeper/higher in the valence/conduction space.


image file: c6ra22249h-f5.tif
Fig. 5 Total and partial density of states of the most energetically favourable configurations at fixed compositions from pure HA (top) to fully-substituted Sn-hydroxylapatite (10 Sn-HA) (bottom) calculated at the GGA/PBE level of theory. The Fermi level is set to zero.

4 Conclusions

The energetics of Sn2+ substitution for Ca2+ in bulk hydroxylapatite (HA) has been investigated within the framework of density functional theory. Calculations of the excess energy, Ex, of Sn-doped HA as a function of the composition predict that compositions of SnxCa10−x(PO4)6(OH)2 (x = 4–9) are predominant, with an optimal stoichiometry of Sn8Ca2(PO4)6(OH)2 (corresponding to a [3 + 3] + 1 Sn(Ia) + 1 Sn(Ib)-trans configuration). At low Sn content, the lowest excess-energy configurations [1 + 0] and [2 + 1]-mer for SnCa9(PO4)6(OH)2 and Sn3Ca7(PO4)6(OH)2, respectively, are not energetically favourable (Ex > 0), while Sn2+ incorporation via coupled substitutions at Ca(II) sites for the [1 + 1]-trans configuration of Sn2Ca8(PO4)6(OH)2 is essential energy neutral. Calculations also show that Sn2+ incorporation at Ca(II) sites is likely to occur up to a composition of Sn6Ca4(PO4)6(OH)2, and further substitutions at Ca(I) sites proceed once full occupancy of Ca(II) sites by Sn2+ is achieved.

Sn substitution into the Ca2+ sublattice of HA results in a steady increase of the unit-cell volume that follows approximately Vegard's law for ideal mixture across the entire compositional range. This volume increase is consistent with predictions from Shannon's rigid ion model accounting for polyhedral distortion and coordination environment. A slight departure from Vegard's volume predictions occurs for x = 0.5 ([3 + 2] configuration) and x = 0.6 ([3 + 3] configuration). Among the predominant compositions (i.e., Ex < 0) of SnxCa1−xHA (x = 0.4–0.9), the volume of the optimal Sn0.8Ca0.2HA ([3 + 3] + 1 Sn(Ia) + 1 Sn(Ib))-trans structure is the closest to Vegard's prediction, thus stressing the importance of the structure–energetics relationship in SnxCa1−xHA solid solutions.

Acknowledgements

Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy's National Nuclear Security Administration under contract DE-AC04-94AL85000. We thank Robert Moore and Mark Rigali (Sandia National Laboratories) for stimulating discussions and John Vienna, Jeff Serne, Matt Asmussen and Nik Qafoku (Pacific Northwest National Laboratory) for useful information.

Notes and references

  1. M. Pasero, A. K. Kampf, C. Ferraris, I. V. Pekov, J. Rakovan and T. J. White, Eur. J. Mineral., 2010, 22, 163 CrossRef CAS.
  2. T. Baikie, S. S. Pramana, C. Ferraris, Y. Huang, E. Kendrick, K. Knight, Z. Ahmad and T. J. White, Acta Crystallogr., Sect. B: Struct. Sci., 2010, 66, 1 CAS.
  3. J. M. Hughes, M. Cameron and K. D. Crowley, Am. Mineral., 1990, 75, 295 CAS.
  4. J. C. Elliott, P. E. Mackie and R. A. Young, Science, 1973, 180, 1055 CrossRef CAS PubMed.
  5. A. Z. Abouzeid, Int. J. Miner. Process., 2008, 85, 59 CrossRef CAS.
  6. J. C. Elliot, Structure and Chemistry of the Apatites and Other Calcium Orthophosphates, Elsevier, Amsterdam, 1994 Search PubMed.
  7. M. Vallet-Regi and J. M. Gonzalez-Calbet, Prog. Solid State Chem., 2004, 32, 1 CrossRef CAS.
  8. F. Mohandes, M. Salavati-Niasari, M. Fathi and Z. Fereshteh, Mater. Sci. Eng., C, 2014, 45, 29 CrossRef CAS PubMed.
  9. H. Zhou and J. Lee, Acta Biomater., 2011, 7, 2769 CrossRef CAS PubMed.
  10. I. Sopyan, M. Mel, S. Ramesh and K. A. Khalid, Sci. Technol. Adv. Mater., 2007, 8, 116 CrossRef CAS.
  11. J. Zeglinski, M. Nolan, M. Bredol, A. Schatte and S. A. M. Tofail, Phys. Chem. Chem. Phys., 2012, 14, 3435 RSC.
  12. T. Tamm and M. Peld, J. Solid State Chem., 2006, 179, 1581 CrossRef CAS.
  13. Y. Tang, H. F. Chappell, M. T. Dove, R. J. Reeder and Y. J. Lee, Biomaterials, 2009, 30, 2864 CrossRef CAS PubMed.
  14. S. Gomes, A. Kaur, J. M. Nedelec and G. Renaudin, J. Mater. Chem. B, 2014, 2, 536 RSC.
  15. K. Matsunaga, H. Murata, T. Mizoguchi and A. Nakahira, Acta Biomater., 2010, 6, 2289 CrossRef CAS PubMed.
  16. S. Gomes, J. M. Nedelec, E. Jallot, D. Sheptyakov and G. Renaudin, Chem. Mater., 2011, 23, 3072 CrossRef CAS.
  17. S. Yin and D. E. Ellis, Phys. Chem. Chem. Phys., 2010, 12, 156 RSC.
  18. J. Xu, T. White, P. Li, C. He and Y. F. Han, J. Am. Chem. Soc., 2010, 132, 13172 CrossRef CAS PubMed.
  19. D. E. Ellis, J. Terra, O. Warschkow, M. Jiang, G. B. Gonzalez, J. S. Okasinski, M. J. Bedzyk, A. M. Rossi and J. G. Eon, Phys. Chem. Chem. Phys., 2006, 8, 967 RSC.
  20. J. Rakovan, R. Reeder, E. V. Elzinga, D. J. Cherniak, C. D. Tait and D. E. Morris, Environ. Sci. Technol., 2002, 36, 3114 CrossRef CAS PubMed.
  21. L. Campayo, A. Grandjean, A. Coulon, R. Delorme, D. Vantelon and D. Laurencin, J. Mater. Chem., 2011, 21, 17609 RSC.
  22. B. M. Thomson, C. L. Smith, R. D. Busch, M. D. Siegel and C. Baldwin, J. Environ. Eng., 2003, 129, 492 CrossRef CAS.
  23. R. C. Moore, Situ formation of apatite for sequestering radionuclides and heavy metals, US Pat., No. 6,592,294, 2003.
  24. R. Klement and H. Haselbeck, Chem. Ber., 1963, 96, 1022 CrossRef CAS.
  25. D. McConnell and D. W. Foreman Jr, Can. Mineral., 1966, 8, 431 CAS.
  26. C. Ganss, A. Lussi, A. Peutzfeldt, N. N. Attia and N. Schlueter, PLoS ONE, 2015, 10, e0123889 Search PubMed.
  27. J. B. Duncan, H. Huber, R. C. Moore, M. J. Rigali, R. B. Malbrouki and E. Brown, Tin(II)apatite: Synthesis, characterization, and challenge with pertechnetate, WM2016 conference, Phoenix, AZ, USA, March 6–10, 2016 Search PubMed.
  28. R. M. Asmussen, J. J. Neeway, A. R. Lawter, T. G. Levitskaia, W. W. Lukens and N. P. Qafoku, J. Nucl. Mater., 2016, 480, 393 CrossRef CAS.
  29. M. W. Billinghurst, D. Jette and E. Somers, International Journal of Radiation Applications and Isotopes, 1981, 32, 559 CrossRef CAS.
  30. M. V. Mikelsons and T. C. Pinkerton, Anal. Chem., 1986, 58, 1007 CrossRef CAS PubMed.
  31. C. Gaillard, C. Den Auwer and S. D. Conradson, Phys. Chem. Chem. Phys., 2002, 4, 2499 RSC.
  32. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
  33. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  34. P. F. Weck and E. Kim, J. Phys. Chem. C, 2016, 120, 16553 CAS.
  35. P. F. Weck, E. E. Kim and E. C. Buck, RSC Adv., 2015, 5, 79090 RSC.
  36. P. F. Weck and E. Kim, Dalton Trans., 2014, 43, 17191 RSC.
  37. K. Kothapalli, E. Kim, T. Kolodziej, P. F. Weck, E. A. Alp, Y. Xiao, P. Chow, C. Kenney-Benson, Y. Meng, S. Tkachev, A. Kozlowski, B. Lavina and Y. Zhao, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 024430 CrossRef.
  38. P. F. Weck, E. Kim, C. F. Jove-Colon and D. C. Sassani, Dalton Trans., 2013, 42, 4570 RSC.
  39. P. F. Weck, E. Kim, C. F. Jove-Colon and D. C. Sassani, Dalton Trans., 2012, 41, 9748 RSC.
  40. B. Lavina, P. Dera, E. Kim, Y. Meng, R. T. Downs, P. F. Weck, S. R. Sutton and Y. Zhao, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 17281 CrossRef CAS PubMed.
  41. M. Corno, A. Rimola, V. Bolis and P. Ugliengo, Phys. Chem. Chem. Phys., 2010, 12, 6309 RSC.
  42. G. Ulian, G. Valdre, M. Corno and P. Ugliengo, American Minerals, 2013, 98, 410 CrossRef CAS.
  43. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  44. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  45. E. R. Davidson, Methods in Computational Molecular Physics, ed. G. H. F. Diercksen and S. Wilson, NATO Advanced Study Institute, Series C, Plenum, New York, 1983, vol. 113, p. 95 Search PubMed.
  46. M. Methfessel and A. T. Paxton, Phys. Rev. B: Condens. Matter Mater. Phys., 1989, 40, 3616 CrossRef CAS.
  47. J. M. Hughes, M. Cameron and K. D. Crowley, Am. Mineral., 1989, 74, 870 CAS.
  48. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Condens. Matter Mater. Phys., 1976, 13, 5188 CrossRef.
  49. W. A. Deer, R. A. Howie and J. Zussman, An introduction to the rock-forming minerals, Longman, UK, 1992 Search PubMed.
  50. N. H. de Leeuw, Chem. Commun., 2001, 1646 RSC.
  51. P. Ptacek, Apatites and their Synthetic Analogues – Synthesis, Structure, Properties and Applications, InTech, 2016 Search PubMed.
  52. F. Monteil Ricera and M. Fedoroff, Sorption of Inorganic Species on Apatites from Aqueous Solutions, in Encyclopedia of Surface and Colloid Science, ed. A. T. Hubbard, Deckker Inc., New York, 2002 Search PubMed.
  53. A. Bigi, M. Gandolfi, M. Gazzano, A. Ripamonti, N. Roveri and S. A. Thomas, J. Chem. Soc., Dalton Trans., 1991, 2883 RSC.
  54. W. Badraoui, A. Bigi, M. Debbabi, M. Gazzano, M. Roverri and R. Thouvenot, Eur. J. Inorg. Chem., 2002, 1864 CrossRef.
  55. L. H. Ahrens, Geochim. Cosmochim. Acta, 1952, 2, 155 CrossRef CAS.
  56. R. D. Shannon and C. T. Prewitt, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1969, 25, 925 CrossRef CAS.
  57. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
  58. J. E. Harries, D. W. L. Hukins and S. S. Hasnain, J. Phys. C: Solid State Phys., 1986, 19, 6859 CrossRef CAS.
  59. F. E. Sowrey, L. J. Skipper, D. M. Pickup, K. O. Drake, Z. Lin, M. E. Smith and R. J. Newport, Phys. Chem. Chem. Phys., 2004, 6, 188 RSC.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.