You are on page 1of 10

Downloaded from http://rspa.royalsocietypublishing.

org/ on December 4, 2016

Proc. R. Soc. A (2011) 467, 2102–2111


doi:10.1098/rspa.2010.0519
Published online 9 February 2011

The initial stages of bioglass dissolution: a


Car–Parrinello molecular-dynamics study
of the glass–water interface
BY ANTONIO TILOCCA1, * AND ALASTAIR N. CORMACK2
1 Department of Chemistry and Thomas Young Centre for Theory and
Simulations of Materials, University College London, 20 Gordon Street,
London WC1H 0AJ, UK
2 New York State College of Ceramics, Alfred University,
Alfred, NY 14802, USA

The initial dissolution stages following implantation of a biomaterial in a physiological


environment are critical for its bioactive properties. Car–Parrinello molecular-dynamics
(CPMD) simulations of the interface between the 45S5 bioglass surface and liquid water
have been carried out to investigate these processes. The analysis of a 40 ps CPMD
trajectory has highlighted the potential mechanism of Na+ /H+ exchange, leading to
formation of surface silanols through water dissociation. Moreover, by comparing the
properties of water layers arranged at different distances from the glass surface, we discuss
the way in which the particular structure and composition of the bioglass surface affects
the hydrogen-bond network and orientation of water in its close proximity.
Keywords: bioglass; molecular-dynamics simulations; dissolution

1. Introduction

When a surface-active bioactive silicate glass is immersed in an aqueous medium,


after implantation of the biomaterial in a physiological environment, a sequence of
inorganic stages leads to the partial dissolution of the glass (Hench 1998; Tilocca
2009). This is a key step towards the bone-bonding and tissue-regeneration ability
of these materials: on one hand, it has been shown that a fast initial dissolution
results in a shorter time needed to form a stable bonding interface with hard
(bone) and soft (muscle) tissues (Peitl et al. 2001); on the other hand, it has
been recently found that some bioactive properties, namely the ‘osteoproductive’
ability to promote growth of new tissue away from the glass surface, depend on the
release of critical amount of ionic species in the surrounding medium, following the
initial dissolution (Xynos et al. 2001). Unlike fully resorbable biomaterials, such
as phosphate glasses (Knowles 2003), bioactive silicate glasses do not completely
dissolve shortly after implant (Hench 1998): the initial dissolution self-passivates
*Author for correspondence (a.tilocca@ucl.ac.uk).
One contribution of 16 to a Special feature ‘High-performance computing in the chemistry and
physics of materials’.

Received 7 October 2010


Accepted 23 November 2010 2102 This journal is © 2011 The Royal Society
Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

Initial stages of bioglass dissolution 2103

the surface against further degradation (Clark et al. 1976), and the presence
of a stable interface is essential to support the deposition and growth of new
tissue. The relevance and the complex nature of the processes occurring at
the glass–liquid interface highlight the need of fundamental investigations of
this region (Tilocca 2010a). Unresolved issues are the nature and strength of
active sites found on the surface, the structure of the hydration layers and their
role in the surface dissolution. Recent Car–Parrinello (CP) simulations (Car &
Parrinello 1985) have started to shed light on these subjects (Tilocca & Cormack
2008, 2009). This computationally demanding ab initio approach, employing
periodic supercells and finite-temperature molecular dynamics (MD), represents
a suitable way to investigate the highly heterogeneous energy landscape of a multi-
component glass surface (Tilocca 2010a). In particular, the strength of selected
surface sites can be probed by examining their interaction with an isolated water
molecule (Tilocca & Cormack 2008), as well as with an extended film mimicking a
bulk solvent (Tilocca & Cormack 2009). These recent simulations have highlighted
the role of exposed Na cations in coordinating water and promoting solvent
penetration in the earliest stages of the dissolution mechanism. Another factor,
also emerged from these simulations, is the unusual kinetic stability of small
silicate rings (compared with those found on the surface of amorphous SiO2 ),
which confirms their potential role as templates for adsorption and growth of the
bone-bonding interface (Sahai & Anseau 2005; Bolis et al. 2008; Tilocca 2010a).
In the present work, we look in more detail at how water molecules are organized
near the surface, focusing on the hydrogen-bond (HB) connectivity and its effect
on the coordination and mechanism of release of Na.

2. Computational methods

CPMD simulations were performed with the QUANTUM ESPRESSO (Giannozzi


et al. 2009) CP code. The electronic structure was treated within the generalized
gradient approximation (GGA) to density-functional theory (DFT), through
the Perdew–Burke–Emzerhof exchange-correlation functional (Perdew et al.
1996). Core-valence electronic interactions were represented through Vanderbilt
ultrasoft pseudopotentials (Vanderbilt 1990), explicitly including Na and Ca
semicore shells in the valence. A plane-wave basis set with cutoffs of 30 and
240 Ry for the smooth part of the wave functions and the augmented charge was
employed, with k-sampling restricted to the G point. A similar computational
set-up has often been employed for modelling water adsorption on oxide surfaces
(Mischler et al. 2005; Adeagbo et al. 2008; He et al. 2009). Using the deuterium
mass for hydrogen allowed us to effectively decouple ionic and electronic degrees
of freedom (Marx & Hütter 2000) with a larger fictitious electronic mass (700
atomic units), which in turn permits a longer time step (0.17 fs).
A slab model of the surface was cut from a bulk sample of the 45S5 bioglass
composition (45.7 SiO2 : 24.3 Na2 O : 27.1 CaO : 2.9 P2 O5 mol%), obtained through
a CPMD melt-and-quench of a periodic cubic box containing 32 SiO2 , 17 Na2 O,
19 CaO and 2 P2 O5 formula units at the experimental density (Tilocca 2007).
A free surface was exposed by doubling one of the sides of the volume-optimized
supercell and keeping three-dimensional periodic boundary conditions on the
resulting orthorhombic cell. A vacuum region thus separates the top face of

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

2104 A. Tilocca and A. N. Cormack

the slab from the periodic image of the bottom face, which is fixed to its bulk-
like geometry. After relaxing the dry slab, we filled the gap with 81 molecules
extracted from a CPMD model of liquid water (with a further small adjustment
of the vertical separation). Now, focusing on the water layers closer to the relaxed
face, the thickness of the gap (approx. 18 Å) ensures that these molecules are only
marginally affected by the periodic image of the opposite (unrelaxed) face: we can
thus assume that the molecules in the first 10 Å above the surface represents a
realistic sample of interfacial liquid water. Whereas small amounts of moisture can
be adsorbed on the as-created bioglass surface before implantation, it is unlikely
that any significant corrosion or degradation of the bioglass surface occurs before
effective contact with a bulk aqueous environment: for instance, the initial step of
the bioactive process, involving Na/proton exchange with the solution, can only
proceed through contact with an aqueous medium. Therefore, assuming that the
properties of the interface are not affected by the presence of a small fraction of
pre-adsorbed water, the dry surface represents a suitable and reproducible model
to study the bioglass–liquid-water interaction.
The fully hydrated system contained 442 atoms and 1952 electrons. A room-
temperature CPMD trajectory of approximately 40 ps was carried out in the
constant number of particles, volume and temperature ensemble using a single
Nosè thermostat, and the first 10 ps were discarded from the analysis.

3. Results and discussion

A snapshot of the glass–water interface, extracted from the CPMD run, is shown
in figure 1. The two most relevant structural features of the surface are already
evident from the picture: (i) the open nature of the bioglass surface, which reflects
the high fragmentation of the bulk silicate network, in turn resulting from the
low-silica content of these materials (Tilocca & Cormack 2007; Tilocca 2010a)
and (ii) the high amount of exposed Na cations, in direct contact with the water
layer (Jallot et al. 2001; Tilocca & Cormack 2009). A more quantitative analysis
is provided by the vertical profile of the number fraction of the different species,
shown in figure 2. The marked increase in the Na fraction above z = 5 Å results
in a net predominance of O (mostly non-bridging oxygens; NBOs), Na and Si
atoms in the top portion of the surface, where a significant water fraction is also
observed. This shows that water can significantly penetrate the as-created bioglass
surface, and this effect seems to be correlated with the Na enrichment in this
region. Favourable H2 O–Na+ and H2 O–Ca2+ interactions on this surface had been
suggested by previous DFT energy optimizations (Tilocca & Cormack 2008): the
calculated adsorption energy of a single water molecule coordinated to Na and Ca
exposed on different surface sites ranged between 0.6 and 1.3 eV. While the broad
range of adsorption energies reflects the widely heterogeneous landscape met by
a water molecule on this surface, the strong Lewis acidity of all exposed network-
modifier cations makes them more favourable as coordination sites than the
siloxane portion of the surface. Putting this effect together with the composition
of the hydrated surface in figure 2, one can conclude that the main water–surface
interaction at the initial interface involves water coordination by NBO− –Na+
pairs, whereas Ca2+ ions have a minor role in the initial stages of surface hydration

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

Initial stages of bioglass dissolution 2105

Figure 1. A snapshot of the bioglass–water interface, extracted from the CPMD trajectory.
The phosphosilicate network (bottom) is represented as yellow/red ball-and-stick, with Na and
Ca cations as green and brown spheres, respectively. Dashed lines highlight HBs.

1.0

0.8
atom fraction (z)

0.6

0.4

0.2

0
–5 0 5 10 15 20 25
z (Å)

Figure 2. Vertical (z) profile of the fraction of atomic species contained in slices 2 Å thick, averaged
over the CPMD trajectory. The centre of the slab model of the surface roughly corresponds to
z = 0. Black circles, glass O; red squares, Si; light green diamonds, P; dark blue triangles, Na;
yellow inverted triangles, Ca; light blue asterisks, water O; violet plus symbols, NBO; dark green
crosses, BO.

and dissolution. Direct inspection of the CPMD trajectory showed that the strong
NBO–water–Na interaction can result in water dissociation with the formation
of a silanol group, according to the reaction
Si−NBO− −Na+ + H2 O = Si−NBO−H + NaOH. (3.1)

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

2106 A. Tilocca and A. N. Cormack

A silanol formed as a result of this process is visible in the snapshot of figure 1.


The resulting OH− is initially associated to the Na cation, but the trajectory
also shows that the OH− charge can be transported elsewhere in the liquid by
a sequence of rapid proton transfers along water molecules (Tilocca & Cormack
2009). On the bioglass surface, the close association with a Lewis acid such as Na+
appears thus necessary to convert an exposed NBO into a silanol group through
water dissociation. As additional Na and eventually Ca cations are transported
from the bulk to the glass surface, they can likewise assist the protonation of
the remaining NBOs: complete protonation of all the initially exposed NBOs can
probably be expected on this basis, even though it is not observed on the relatively
short 40 ps time scale of the CPMD run. The overall effect of reaction (3.1)
is the Na+ /H+ exchange between the glass and the solution, resulting in
increased local alkalinity, which is in turn considered important to promote
further degradation of the glass network through breaking of Si−O−Si bridges
(Hench 1998).
The open nature of the 45S5 surface somewhat complicates the exact
identification of NBOs, which will be protonated, because in this system, water
molecules can, in principle, reach and dissociate at NBOs located well below
the top of the surface (Tilocca 2010a). The extent of this penetration cannot
be determined a priori, and therefore, the number of protonated NBOs must be
estimated based on an arbitrary cutoff separating ‘active’ and ‘inactive’ NBOs.
Based on figure 2, denoting water penetration at least in the surface region above
z = 5 Å, the total number of NBOs found in this region provides a lower limit of
approximately 4 nm−2 to the final surface density of silanol groups in the fully
hydrated system.
The effect of the glass surface on the structural properties of the hydrated water
can be examined by comparing the structure of the water molecules in direct
contact with the surface (‘surface water’; SW) and those found at higher vertical
distances from the surface (‘bulk water’; BW). The z-profiles in figure 2 allow us
to assign water molecules within the [5 Å < z < 10 Å] layer as SW, and those with
[10 Å < z < 15 Å] as BW; in other words, SW molecules access regions where also
atoms belonging to the glass can be found, whereas no glass atoms are present in
the BW layer. The water–water radial distribution functions (rdfs) for SW and
BW are shown in figure 3. When comparing the rdfs calculated in the two regions,
one should take into account the non-spherical and inhomogeneous distribution
of water surrounding molecules in a thin rectangular slice: for instance, no BW
is found below the SW layer; therefore, the average number of water molecules
resident in each layer and the corresponding limit homogeneous density were used
to normalize the rdfs in figure 3.
Notwithstanding the tendency of the adopted computational framework to
produce over-structured rdfs and overestimate the number of HBs in room-
temperature liquid water (Sit & Marzari 2005; Todorova et al. 2006), the relative
trends examined here can be considered accurate and representative. The SW
and BW rdfs are rather similar, with the intermolecular O−O and O−H peaks
of the SW at slightly shorter distances, denoting stronger HBs involving these
molecules. This small strengthening of water–water HBs close to the surface can
be further investigated by separately calculating the HB statistics of SW and BW.
A geometrical definition of HB was used (Tilocca & Selloni 2004a): two molecules
were considered linked by an HB when their O−O distance was less than 3.3 Å,
Proc. R. Soc. A (2011)
Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

Initial stages of bioglass dissolution 2107

Table 1. Hydrogen-bond statistics. nHB is the average number of hydrogen bonds per molecule;
< ROO > and < ROH > (Å) are the mean O−O and intermolecular O−H length of H-bonded pairs.
The SW and BW columns show the overall HB statistics for SW and BW molecules; for water–water
interactions, these are further decomposed to analyse HBs between two SW molecules (SW−SW),
an SW and a BW molecule (SW−BW), and two BW molecules (BW−BW). Standard errors are
quoted in parentheses.

SW SW−SW SW−BW BW−BWa BW

water–water
nHB 1.88 (0.09) 1.25 (0.1) 0.63 (0.07) 2.48 (0.03) 3.56 (0.04)
< ROO > (Å) 2.80 2.79 2.84 2.83 2.83
< ROH > (Å) 1.83 1.81 1.87 1.85 1.85
water–surface
nHB 0.59 (0.03) —
< ROO > (Å) 2.80 —
< ROH > (Å) 1.83 —
an
HB (BW) in the last column also includes HBs between a BW molecule and molecules located
above the BW slice, hence nHB (BW) > [nHB (SW−BW) + nHB (BW−BW)].

the OH intermolecular distance was less than 2.4 Å and the OOH angle between
the O–O vector and the covalent OH bond was less than 30◦ . The results of the HB
analysis are reported in table 1. Whereas a molecule in the BW region is H-bonded
to 3.6 other water molecules on average, the close proximity of the surface reduces
the number of water–water HBs for SW molecules by about a half. While some of
the lost water–water HBs are balanced by analogous water–surface interactions,
the total number of HBs formed by an SW molecule is significantly lower than
BW: only approximately 2.5 HBs/molecule in total, including 0.6 HBs/molecule
formed with the glass surface. The previous discussion suggests that residual
interactions (assuming a pseudo-tetrahedral water coordination) are represented
by Na+ −O(H2 ) coordination; as a matter of fact, the OSW −Na+ coordination
number, calculated by integrating the corresponding rdfs (not shown) up to
the first minimum, is 0.8. Therefore, the HB statistics show that the reduced
water–water coordination in the surface region is partially balanced by HBs with
surface NBO (or NBO–H), and by coordination to Na+ . Furthermore, table
1 shows that the small strengthening of HBs in the surface region, inferred
from the rdfs, originates from individual HBs between two SW being stronger
than those between an SW and a BW, and than those between two BWs.
Whereas the interaction with the surface atoms probably plays a role in it, further
investigations (for instance of the electronic-charge rearrangements induced upon
water contact with the surface) are needed to identify the source of this effect.
The reduced availability of HB partners in the surface region has other
structural consequences for the water molecules located there: the distribution
of the angle q formed by the molecular bisector and the surface normal (z-axis) is
shown in figure 4 for SW and BW molecules. SW is oriented with the molecular
dipole roughly aligned (with an approx. 30◦ tilt) to the surface normal: the
parallel alignment, with the molecular dipole pointing up (away from the surface)

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

2108 A. Tilocca and A. N. Cormack

3
g (r)
2

0
2 3 4 5 6 7
r (Å)

Figure 3. Water–water radial distribution functions (rdfs) for surface water (SW) and bulk water
(BW). The rdfs normalization takes into account the different water density around the SW and
BW slabs. Black continuous line, O−O (SW); red continuous line, O−H (SW); black dashed line,
O−O (BW); red dashed line, O−H (BW).

0.06

0.04
f (θ)

0.02

0 5 100 150
angle θ (°)

Figure 4. Distribution of the angle q between the water dipole (bisecting the HOH angle) and the
surface normal, for SW and BW molecules. Black solid line, SW; red solid line, BW.

is preferred, whereas the secondary peak at 150◦ arises from water molecules
with their dipole pointing down, towards the surface. On the other hand, the
most common orientation for BW molecules is ‘complementary’ to the SW ones,
with their molecular dipole roughly perpendicular to the surface normal, and a
net corresponding decrease in the probability of orientations with the molecular
dipole pointing up or down. This pattern can be interpreted on the basis of the
reduced availability of HB partners (acceptors) for the SW molecules, which forces
them to direct one hydrogen upward (towards the BW molecules) or downward

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

Initial stages of bioglass dissolution 2109

(towards surface NBOs). Because the oxygen of dipole-up SW molecules is


frequently coordinated to an Na cation, these orientations are favoured by the
initial abundance of Na cations in the glass surface. A BW molecule in the region
above SW can either accept an HB from the dipole-up SW, or donate an HB to
the dipole-down SW, and figure 4 shows that these configurations involve q angles
around 90◦ . Therefore, even though more HB partners surround BW molecules
(table 1), the structured distribution of the orientation of the SW partners induces
a corresponding ‘complementary’ anisotropy also in the BW distribution.

4. Conclusions

The high NBO : BO ratio of the bulk 45S5 bioglass composition, with its low SiO2
content, results in a marked fragmentation of the silicate network. This leads to a
fast initial release of ionic dissolution products upon contact with a physiological
(aqueous) medium, which is one of the keys to the bioactive behaviour of these
materials. The present ab initio simulations provide a high-resolution view into
the structure and reactivity of the bioglass–liquid-water interface.
We have shown that the initial Na+ /H+ exchange at the interface involves
water dissociation at an Na+ –NBO− pair, with release of the Na cation
from the surface: because transport of further Na+ ions from the bulk glass
to the surface is thus essential to promote complete protonation of surface
NBOs, additional CPMD simulations are currently addressing Na migration
in these materials (Tilocca 2010b). Moreover, the effect of surface interactions
on the water structure and connectivity has been investigated; the relatively
low availability of favourable targets for HB interactions on the surface
significantly affects water coordination in this region. Na cations partially enter
the coordination sphere of an SW, and the reduced density of suitable HB
acceptors also determines the preferential upward or downward orientation
of water dipoles observed in the surface region, which in turn appears to
influence the orientation of the water molecules immediately above them. It is
interesting to note how, in spite of the low density and disordered nature of
the 45S5 surface, its effect can still be transmitted to molecules not in direct
contact with it, as was found for dense crystalline oxide surfaces (Tilocca &
Selloni 2004b). Moreover, we found that whereas the overall HB connectivity
decreases in the surface region, the strength of individual HBs increases: further
calculations are needed to address the electronic and dynamical causes of
these effects.
A.T. thanks the UK’s Royal Society for support (RS-URF). Computer resources on the HPCx and
HECToR national supercomputing services were provided via the UK’s HPC Materials Chemistry
Consortium, funded by EPSRC (grants EP/D504872 and EP/F067496).

References
Adeagbo, W. A., Doltsinis, N. L., Klevakina, K. & Renner, J. 2008 Transport processes at
a-quartz–water interfaces: insights from first-principles molecular dynamics simulations. Chem.
Phys. Chem. 9, 994–1002. (doi:10.1002/cphc.200700819)

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

2110 A. Tilocca and A. N. Cormack

Bolis, V., Busco, C., Aina, V., Morterra, C. & Ugliengo, P. 2008 Surface properties of silica-based
biomaterials: Ca species at the surface of amorphous silica as model sites. J. Phys. Chem. C
112, 16 879–16 892. (doi:10.1021/jp805206z)
Car, R. & Parrinello, M. 1985 Unified approach for molecular dynamics and density-functional
theory. Phys. Rev. Lett. 55, 2471–2474. (doi:10.1103/PhysRevLett.55.2471)
Clark Jr, A. E., Pantano Jr, C. J. & Hench, L. L. 1976 Auger spectroscopic analysis
of bioglass corrosion films. J. Am. Ceram. Soc. 59, 37–39. (doi:10.1111/j.1151-2916.1976.
tb09382.x)
Giannozzi, P. et al. 2009 QUANTUM ESPRESSO: a modular and open-source software project for
quantum simulations of materials. J. Phys., Condens. Matter 21, 395502. (doi:10.1088/0953-
8984/21/39/395502)
He, Y., Tilocca, A., Dulub, O., Selloni, A. & Diebold, U. 2009 Local ordering and electronic
signatures of submonolayer water on anatase TiO2 (101). Nat. Mat. 8, 585–589. (doi:10.1038/
nmat2466)
Hench, L. L. 1998 Bioceramics. J. Am. Ceram. Soc. 81, 1705–1728. (doi:10.1111/j.1151-2916.1998.
tb02540.x)
Jallot, E., Benhayoune, H., Kilian, L., Josset, Y. & Balossier, G. 2001 An original method to assess
short-term physicochemical reactions at the periphery of bioactive glass particles in biological
fluids. Langmuir 17, 4467–4470. (doi:10.1021/la001669m)
Knowles, J. C. 2003 Phosphate based glasses for biomedical applications. J. Mater. Chem. 13,
2395–2401. (doi:10.1039/b307119g)
Marx, D. & Hütter, J. 2000 Ab initio molecular dynamics: theory and implementation. In Modern
methods and algorithms of quantum chemistry, NIC series, vol. 1 (ed. J. Grotendorst), pp.
301–449. Jülich, Germany: John von Neumann Institute for Computing.
Mischler, C., Horbach, J., Kob, W. & Binder, K. 2005 Water adsorption on amorphous silica
surfaces: a Car–Parrinello simulation study. J. Phys., Condens. Matter 17, 4005. (doi:10.1088/
0953-8984/17/26/001)
Peitl, O., Zanotto, E. D. & Hench, L. L. 2001 Highly bioactive P2 O5 –Na2 O–CaO–SiO2 glass-
ceramics. J. Non-Cryst. Solids 292, 115–126. (doi:10.1016/S0022-3093(01)00822-5)
Perdew, J. P., Burke, K. & Ernzerhof, M. 1996 Generalized gradient approximation made simple.
Phys. Rev. Lett. 77, 3865–3868. (doi:10.1103/PhysRevLett.77.3865)
Sahai, N. & Anseau, M. 2005 Cyclic silicate active site and stereochemical match for apatite
nucleation on pseudowollastonite bioceramics-bone interfaces. Biomaterials 26, 5763–5770.
(doi:10.1016/j.biomaterials.2005.02.037)
Sit, P. H.-L. & Marzari, N. 2005 Static and dynamical properties of heavy water at ambient
conditions from first-principles molecular dynamics. J. Chem. Phys. 122, 204510. (doi:10.1063/
1.1908913)
Tilocca, A. 2007 Structure and dynamics of bioactive phosphosilicate glasses and melts from
ab initio molecular dynamics simulations. Phys. Rev. B 76, 224202. (doi:10.1103/PhysRevB.
76.224202)
Tilocca, A. 2009 Structural models of bioactive glasses from molecular dynamics simulations. Proc.
R. Soc. A 465, 1003–1027. (doi:10.1098/rspa.2008.0462)
Tilocca, A. 2010a Models of structure, dynamics and reactivity of bioglasses: a review. J. Mater.
Chem. 20, 6848–6858. (doi:10.1039/c0jm01081b)
Tilocca, A. 2010b Sodium migration pathways in multicomponent silicate glasses: Car–Parrinello
molecular dynamics simulations. J. Chem. Phys. 133, 014701. (doi:10.1063/1.3456712)
Tilocca, A. & Cormack, A. N. 2007 Structural effects of phosphorus inclusion in bioactive silicate
glasses. J. Phys. Chem. B 111, 14 256–14 264. (doi:10.1021/jp075677o)
Tilocca, A. & Selloni, A. 2004a Structure and reactivity of water layers on defect-free and defective
anatase TiO2 (101) surfaces. J. Phys. Chem. B 108, 4743–4751. (doi:10.1021/jp037685k)
Tilocca, A. & Selloni, A. 2004b Vertical and lateral order in adsorbed water layers on anatase
TiO2 (101). Langmuir 20, 8379–8384. (doi:10.1021/la048937r)
Tilocca, A. & Cormack, A. N. 2008 Exploring the surface of bioactive glasses: water adsorption
and reactivity. J. Phys. Chem. C 112, 11 936–11 945. (doi:10.1021/jp803541j)

Proc. R. Soc. A (2011)


Downloaded from http://rspa.royalsocietypublishing.org/ on December 4, 2016

Initial stages of bioglass dissolution 2111

Tilocca, A. & Cormack, A. N. 2009 Modeling the water–bioglass interface by ab-initio molecular
dynamics simulations. ACS Appl. Mater. Interfaces 1, 1324–1333. (doi:10.1021/am900198t)
Todorova, T., Seitsonen, A. P., Hutter, J., Kuo, I. W. & Mundy, C. J. 2006 Molecular dynamics
simulation of liquid water: hybrid density functionals. J. Phys. Chem. B 110, 3685–3691.
(doi:10.1021/jp055127v)
Vanderbilt, D. 1990 Soft self-consistent pseudopotentials in a generalized eigenvalue formalism.
Phys. Rev. B 41, 7892–7895. (doi:10.1103/PhysRevB.41.7892)
Xynos, I. D., Edgar, A. J., Buttery, L. D. K., Hench, L. L. & Polak, J. M. 2001 Gene-expression
profiling of human osteoblasts following treatment with the ionic products of bioglass(R) 45S5
dissolution. J. Biomed. Mater. Res. 55, 151–157. (doi:10.1002/1097-4636(200105)55:2<151::
AID-JBM1001>3.0.CO;2-D)

Proc. R. Soc. A (2011)

You might also like