You are on page 1of 9

A computational study of the surface structure and reactivity of

calcium fluoride

Nora H. de Leeuw*a,b and Timothy G. Cooperb,c


a
School of Crystallography, Birkbeck College, University of London, Malet Street, London,
UK WC1E 7HX. E-mail: n.deleeuw@mail.cryst.bbk.ac.uk
b
Department of Chemistry, University College, 20 Gordon Street, London, UK WC1H 0AJ
c
Department of Chemistry, University of Reading, Whiteknights, Reading, UK RG6 6AD

Received 15th August 2002, Accepted 2nd October 2002


First published as an Advance Article on the web 12th November 2002

Electronic structure calculations based on the density functional theory (DFT) are employed to investigate the
electronic structure of fluorite (CaF2) and the mode and energies of adsorption of water at the main {111}
cleavage plane. Electron density plots show the crystal to be strongly ionic with negligible ionic relaxation of
the unhydrated surface. We find associative adsorption of water at the surface with hydration energies between
41 and 53 kJ mol21, depending on coverage. We next employ atomistic simulation techniques to investigate the
competitive adsorption of water and methanoic acid at the planar and stepped {111}, {011} and {310}
surfaces. The hydration energies and geometries of adsorbed water molecules on the planar {111} surface agree
well with those found by the DFT calculations, validating the interatomic potential parameters. Methanoic acid
adsorbs in completely different configurations on the three surfaces, but always by one or both oxygen atoms
to one or more surface calcium atoms. Molecular Dynamics simulations at 300 K show that the effect of
temperature is to increase the difference in adsorption energy between methanoic acid and water at the planar
{111} surface. The methanoic acid remains bound to the surface whereas the water molecules prefer to form a
droplet of water between the two surface planes. We show in a series of calculations of the co-adsorption of
water and methanoic acid that the presence of solvent makes a significant contribution to the final adsorption
energies and that the explicit inclusion of solvent in the calculations is necessary to correctly predict relative
reactivities of different surface sites, a finding which is important in the modelling of mineral separation
processes such as flotation.

1 Introduction potential techniques to model larger scale systems. Comparison


of the geometries and energies of adsorbed water molecules at
Fluorite (CaF2) is a widely distributed mineral and it often the calcium fluoride surface, obtained using both quantum
occurs in combination with a range of other ores, notably lead mechanical and classical techniques, will give an indication as
and tin ores, and together with calcite and apatites.1 It is the to whether the potential model used in the atomistic simu-
most important source material for hydrofluoric acid2 and as lations is capable of reliably modelling fluorite–water interac-
such needs to be efficiently separated from any co-existing tions. If there is good agreement between the results of the
minerals, usually by the technique of froth flotation, e.g. refs. 3 different methods, we can then be confident of applying ato-
and 4. In addition to numerous investigations of the fluorite mistic simulations to larger systems, including both adsorbing
structure itself, both experimental and computational, e.g., species, which are currently beyond the capability of electronic
refs. 5–7 its crystal growth and dissolution have been widely structure calculations.
studied, e.g., refs. 8–10, as well as its possible application as a
substrate for the epitaxial growth of thin films.11 The effect of
hydration on the CaF2 surface structure is of considerable
2 Methodology
interest in all of these applications, while the adsorption of
organic surfactant molecules at the surfaces is important in In the electronic structure calculations, the total energy and
mineral processing techniques, such as flotation, where the structure of the simulation system, comprising slabs of solid
selectivity of the surfactants plays a major rôle in the design of material separated by a vacuum gap, which together are
a successful separation process.3,4 repeated periodically in three dimensions, was determined
In this paper, we describe our computational investigations using the Vienna Ab Initio Simulation Program (VASP).12–15
of the electronic structure of the fluorite crystal and the sur- The basic concepts of density functional theory (DFT) and the
face reactivity towards water and methanoic acid. We use principles of applying DFT to pseudopotential plane-wave
methanoic acid as a model of carboxylic acid surfactants such calculations has been extensively reviewed elsewhere.16–18
as oleic acid, and study its adsorption at the major {111}, {011} Furthermore, this methodology is well established and has
and {310} surfaces of calcium fluoride, including a series of been successfully applied to the study of adsorbed atoms and
stepped surface sites. The approach we have chosen to adopt is molecules on the surface of ionic materials.19–22 The VASP
to use electronic structure calculations based on the density program employs ultra-soft pseudo-potentials,23,24 which
functional theory (DFT) to study the main {111} cleavage allows a smaller basis set for a given accuracy. Within the
plane of fluorite in order to obtain, firstly, details of the pseudo-potential approach only the valence electrons are
geometry and electronic structure of the dry and hydrated treated explicitly and the pseudo-potential represents the
surfaces, and secondly, reliable estimates of the hydration effective interaction of the valence electrons with the atomic
energies to compare with when using classical interatomic cores. In our calculations the core consisted of orbitals up to

DOI: 10.1039/b208004d J. Mater. Chem., 2003, 13, 93–101 93


This journal is # The Royal Society of Chemistry 2003
and including the 1s orbital for fluorine and oxygen and the 3p fluoride ions both above and below.9 The {111} surface is thus
orbital for Ca (H has no core). The valence orbitals are terminated with fluorine atoms and just below the surface are
represented by a plane-wave basis set, in which the energy of seven-coordinate calcium ions. We first employed DFT
the plane-waves is less than a given cutoff (Ecut). methods to investigate the dehydrated {111} surface, calculat-
For surface calculations, where two energies are compared, it ing lattice parameters of a ~ b ~ c ~ 5.4051 Å, in excellent
is important that the total energies are well converged. The agreement with experiment. Fig. 2 shows the relaxed {111}
degree of convergence depends on a number of factors, two of surface, including the electron density distribution around the
which are the plane-wave cutoff and the density of k-point calcium and fluoride ions and interatomic distances, from
sampling within the Brillouin zone. We have by means of a which it is clear that the ionic relaxation of the surface is small,
series of test calculations on bulk CaF2, where these parameters a dilation of the topmost F–Ca spacing of 0.01 Å, followed by a
were varied systematically, determined values for Ecut (500 eV) contraction of the second interlayer distance by 0.02 Å. The
and the size of the Monkhorst–Pack25 k-point mesh (3 6 3 6 3) contour plots of the electron density show that the crystal is
so that the total energy is converged to within 0.05 eV. strongly ionic with electron density firmly centred on the
The electronic structure calculations were performed within anions. The distortion of the electron density round the surface
the generalized-gradient approximation (GGA), using the ions is minimal, which leads to the negligible ionic relaxation of
exchange-correlation potential developed by Perdew and the surface layer.
Wang,26 which approach has been shown to give reliable
results for the energetics of adsorbates, e.g., water on CaO,27
3.1 Hydrated {111} surface
TiO2 and SnO2.28
The larger scale systems were modelled using the less We next investigated the adsorption of water at the {111}
computationally expensive atomistic simulation techniques, surface to evaluate the energies of adsorption and the relaxed
based on the Born model of solids,29 where simple para- hydrated surface structure. We studied both the adsorption of a
meterised analytical forms are used to describe the forces full monolayer (i.e. one water molecule per surface calcium ion)
between atoms. In this work we employ the METADISE and a 50% partial coverage. We used a range of different
code30 to investigate the surface systems by energy minimisa- starting configurations of associatively adsorbed water mole-
tion, which is achieved by adjusting the atoms in the system cules on the surface to ensure that the final converged
until the net forces on each atom are zero. Energy minimisation configuration would be a global, rather than a local, minimum
simulations will yield adsorption energies, which have pre- energy configuration. The hydration energy per water mole-
viously been shown to give good agreement with experimental cule for the partial coverage of 50% was calculated to be
surface sampling techniques such as temperature programmed 253.4 kJ mol21 compared to the sum of the energies for the
desorption, e.g., ref. 31, as well as lowest energy configurations dry surface and an isolated gaseous water molecule, which
of the adsorbate/solid interface. decreased to 241.4 kJ mol21 for full monolayer coverage.
In addition, we employed Molecular Dynamics (MD) These calculated hydration energies of approximately 41–
simulations to derive the potential parameters for the water– 53 kJ mol21 suggest that the water molecules are physisorbed
methanoic acid interactions and also to investigate whether the rather than chemisorbed onto the surface. At 50% partial
inclusion of temperature in the calculations would affect the coverage (Fig. 3), the water molecules adsorb almost flatly
adsorption behaviour and/or energies. The MD code used was onto the surface and are much more closely coordinated to
DL_POLY32 where the integration algorithms are based the surface than at full monolayer coverage; the calcium–
around the Verlet leap-frog scheme.33 We used the Nosé– oxygen distance increases from 2.37 Å to 2.62 Å and the
Hoover algorithm for the thermostat,34,35 as this algorithm fluorine–hydrogen distance from 1.52 Å to 1.67 Å when the
generates trajectories in both NVT and NPT ensembles, thus coverage is increased. This result suggests that the lattice
keeping our simulations consistent. The Nosé–Hoover para- spacing of fluorine is not large enough to accommodate a full
meters were set at 0.5 ps for both the thermostat and barostat layer of water molecules in an optimum position.
relaxation times. The surface simulations were run for at least The DFT calculations did not show any dissociation of the
500 ps each (approximately 2.5 6 106 timesteps) as an NVT water molecules to form a hydroxylated surface at either
ensemble, i.e., a constant number of particles, constant volume coverage, so in order to be certain that there was no lower
and a constant temperature of 300 K. energy configuration with dissociatively adsorbed water
We used a combination of three potential models for a molecules, we also simulated a fully hydroxylated surface as
description of the interactions of the various atoms in the a starting configuration. A hydroxyl group was placed above
systems, namely by Catlow et al., for the calcium fluoride each surface calcium ion and a proton above each surface
crystal;36 the cvff forcefield for methanoic acid;37 and the water fluoride ion. However, the dissociatively adsorbed water
potential model by de Leeuw and Parker.38 The parameters for molecules reassembled to form molecular water. Fig. 4 shows
the interactions between water and methanoic acid with the the sequence of reformation of the water molecules on the
fluoride surfaces were derived following the approach by {111} surface, from initial configuration to a midway snapshot,
Schröder et al.,39 while the water–methanoic acid parameters where tilting of the hydroxyl group towards the proton and
were fitted to the experimental solvation energy of methanoic lengthening of the H–F bond (from 1.2 Å to 1.7 Å) is observed,
acid.40 The full potential model is given in Table 1. to the final configuration with associatively adsorbed water
molecules. The distance between the hydroxyl oxygen atom
and the proton decreases from an initial hydrogen-bond
3. Results and discussion distance of 2.22 Å to a normal O–H bond of 1.01 Å. The
calculated hydration energy of 241.3 kJ mol21 is identical to
Calcium fluoride has the cubic fluorite crystal structure with the associative starting configuration (monolayer coverage),
space group Fm3m and a ~ b ~ c ~ 5.4323 Å, where each giving us confidence that the lowest energy configuration had
calcium ion is surrounded by eight fluoride ions, which are in been found. The easy reformation of undissociated water
turn coordinated to four calcium ions in a tetrahedral molecules indicates that there is no significant energy barrier to
arrangement, shown in Fig. 1. The calcium ions are arranged this process.
on a cubic face-centred lattice, and if we divide the unit cell into The preference for associatively rather than dissociatively
8 smaller cubes, we find the fluoride ions in the centres of these adsorbed water on the main CaF2 {111} surface agrees
cubes.41 The cleavage plane is the {111} surface, which consists qualitatively with previous atomistic and electronic structure
of planes of calcium ions in a hexagonal array with a layer of calculations of water adsorption at the vacuum interface of

94 J. Mater. Chem., 2003, 13, 93–101


Table 1 Potential parameters used in this work (short range cutoff 20 Å)

Charges (e)

Ion Core Shell Core-shell interaction/eV Å22

F 11.380 22.380 101.200000


Oxygen of carbonate group (O) 10.587 21.632 507.400000
Oxygen of water (Ow) 11.250 22.050 209.449602
Ca 12.000
Carbon of carbonate group (C) 11.135
Hydrogen of water (Hw) 10.400
Doubly-bonded oxygen of methanoic acid (OD) 20.380
Hydroxy oxygen of methanoic acid (OH) 20.380
Carbon of methanoic acid (CD) 10.310
Hydroxy hydrogen of methanoic acid (HO) 10.350
Hydrogen attached to carbon of methanoic acid (HC) 10.100
Buckingham potential
Ion pair A/eV r/Å C/eV Å6
Ca–O 1550.0 0.29700 0.0
Ca–F 1272.8 0.2997 0.0
Ca–Ow 1186.6 0.29700 0.0
Hw–O 396.3 0.23000 0.0
Hw–Ow 396.3 0.25000 10.0
O–O 16372.0 0.21300 3.47
F–F 99731833.99084 0.12013 17.02423
O–Ow 12533.6 0.21300 12.09
Ca–OH 563.64 0.29700 0.0
F–Ow 79785220.99 0.12013 26.78752
F–Hw 715.339 0.2500 10.00
Ca–OD 563.64 0.29700 0.0
OH–O 37898119 0.12013 11.309
OD–O 37898119 0.12013 11.309
OH–F 37898119 0.12013 25.1
OD–F 37898119 0.12013 25.1
Ow–OH 4797.6 0.213 30.2
Ow–OD 4797.6 0.213 30.2
Ow–HO 396.3 0.25 0.0
Ow–HC 396.3 0.25 0.0
Ow–CD 895 0.26 0.0
Lennard–Jones potential
A/eV Å12 B/eV Å6
Ow–Ow 39344.98 42.15
HC–O 2915.25 4.222
HO–O 2915.25 4.222
CD–O 3315.91 19.846
OD–Hw 1908.1 5.55
OH–Hw 1908.1 5.55
HC–F 2915.25 9.3784
HO–F 2915.25 9.3784
CD–F 3315.91 44.012
Morse potential
D/eV a/Å21 r0/Å
C–O 4.710000 3.80000 1.18000
Hw–Ow 6.203713 2.22003 0.92376
CD–HC 4.66 1.77 1.10
OH–HO 4.08 2.28 0.96
CD–OH 4.29 2.00 1.37
CD–OD 6.22 2.06 1.23
Three-body potential
k/eV rad22 H0̃
Ocore–C–Ocore 1.69000 120.000000
H–Owshell–H 4.19978 108.693195
OH–HO–CD 4.29 112.000000
CD–HC–OH 4.72 110.000000
CD–OD–HC 4.72 120.000000
CD–OD–OH 12.45 123.000000
Four-body potential
k/eV rad22 H0̃
C–Ocore–Ocore–Ocore 0.11290 180.0
Intermolecular Coulombic interaction (%)
Hw–Ow 50
Hw–Hw 50

MgO, another ionic crystal of cubic space group Fm3m, which and only occurs at defects and low-coordinated surface sites44–
46
showed that dissociative adsorption is energetically unfavour- or at the liquid water interface where H3O1 species are taken
able on the perfect {100} cleavage plane, e.g., refs. 19,31,42,43, into account.43 Electronic structure calculations of the main

J. Mater. Chem., 2003, 13, 93–101 95


Fig. 3 Plan view of the minimum energy structure of the CaF2 {111}
surface with 50% coverage of associatively adsorbed water molecules,
showing almost flat adsorption of the water molecules (fluorite shown
as framework, water space-filled; Ca ~ black, F ~ pale grey, O ~
black, H ~ white).

Fig. 1 Bulk structure of CaF2 showing cubic face-centred calcium


lattice with the fluoride ions in the centres of each of eight smaller cubes
making up the cubic unit cell (Ca ~ black, F ~ pale grey).

Fig. 2 Side view of the relaxed CaF2 {111} surface showing electron
density contour plots and interatomic distances (Ca ~ dark grey, F ~
pale grey, contour levels are from 0.05 to 0.35 e Å23 at 0.05 e Å23
intervals, bond lengths in Å).

TiO2 cleavage plane, the {110} surface, show hydroxylation of


the surface at half coverage.20,28 However, Lindan et al. found
that at full coverage a mixture of associatively and dissocia-
tively adsorbed water molecules is observed, where the water Fig. 4 Reassembly of dissociatively adsorbed water molecules on the
molecule is adsorbed almost flat onto the surface to maximise CaF2 {111} surface: (a) side view showing initial configuration of
hydrogen-bonding to oxygen atoms of both the hydroxyl group hydroxylated surface; (b) snapshot during minimisation process
showing tilting of hydroxyl group and lengthening of F–H bond; (c)
and the mineral surface.20 This configuration of the associa- side view of the final configuration showing associatively adsorbed
tively adsorbed water molecules on the TiO2 {110} surface is water molecules (crystal shown as framework, water space-filled; Ca ~
thus like that on the CaF2 {111} surface, where almost flat black, F ~ pale grey, O ~ black, H ~ white).

96 J. Mater. Chem., 2003, 13, 93–101


adsorption of the water molecules and a network of hydrogen- with close coordination of both hydrogens to surface fluoride
bonding is preferred over dissociative adsorption. ions, while the water molecules in the DFT calculations adsorb
Our calculations indicate that the binding of the water slightly tilted, with one hydrogen atom pointing away from the
molecule’s oxygen atom to a surface calcium atom is the main surface, giving very different H–F distances for the two hydro-
interaction. This finding suggests that increasing the coordina- gens (1.52 Å and 2.85 Å). When a full monolayer is adsorbed,
tion of the surface cation to the bulk value, from seven- to the average adsorption energy drops to 38.5 kJ mol21,
eight-coordinate for the {111} surface, is the driving force compared to 41.4 kJ mol21 for the DFT calculations. Due to
behind the adsorption. From these calculations we would lack of space on the surface for a full monolayer, the water
therefore suggest that, as with MgO, dissociative adsorption of molecules do not adsorb flat onto the surface and the lesser
water takes place at defects and low-coordinated surface sites binding between surface fluoride ions and hydrogen atoms
rather than the higher-coordinated cations of the perfect {111} leads to an even better agreement for the adsorption energies
cleavage plane. between the two computational techniques. On the {011} and
{310} surfaces, the Ca–Ca spacings are large enough easily to
accommodate a full monolayer of water. On the {011} surface,
3.2 Hydration of planar and stepped surfaces
the water molecules adsorb in an upright fashion, without
In order to study larger-scale systems, we employed atomistic significant H–F interactions. However, the increased stability
simulation techniques to model two more fluorite surfaces in of the hydrated {310} surface is due to flat adsorption of the
addition to the {111} surface, namely the {011} and {310} water molecules, similar to the {111} surface, and an extensive
surfaces, as well as two stepped {111} surfaces, as a more network of hydrogen-bonding between both hydrogens and
realistic model for experimental surfaces. We investigated surface fluoride ions.
adsorption of both water and methanoic acid on the surfaces to As ‘real’ surfaces are never completely free from defects, we
evaluate their relative structures and adsorption energies. have also included stepped surface sites in our calculations. We
We first modelled the unhydrated surfaces to calculate their considered two steps on the {111} surface that differ in the
stabilities, after which we hydrated the surfaces to evaluate any orientation of the F2 groups, which either lean backwards at an
changes in stability. The surface stabilities are measured by the obtuse angle of 135u with respect to the underlying plane or
surface energy, which is calculated as follows: forwards at an acute angle of 45u (Fig. 5). From the adsorption
energies in Table 3, we see that hydration of the acute step edge
Us {Ub is less favourable than the planar surface, due to the restricted
c~
A space available for the adsorbing water molecule under the
step. However, the more open adsorption site at the obtuse step
where Us is the energy of the ions in the surface simulation cell edge, combined with the step ions’ lower coordination number,
and Ub is the energy of an equal number of bulk ions, while A is makes hydration of this step more exothermic than the planar
the area of the surface. A low, positive value for the surface surface.
energy indicates a stable surface, which will be important in the
morphology of the mineral. The surface energies of the dry and
hydrated surfaces are collected in Table 2, where the surface 3.3 Adsorption of methanoic acid
energy of the hydrated surface is calculated with respect to bulk
water. The {111} surface is clearly the most stable of the three When we considered adsorption of methanoic acid at the same
surfaces considered, both in dry and aqueous conditions, in surface sites, we found that the methanoic acid molecules
agreement with the fact that this is experimentally the perfect adsorb onto the planar surfaces in three distinctly different
cleavage plane of calcium fluoride.41 The dry {310} surface is
very unstable, but its stability is increased substantially when
the surface is hydrated, making it now more stable than the
{011} surface.
We first considered adsorption of water on the planar {111}
surface, where comparison to the equivalent DFT calculations
gives an indication of the accuracy of the potential model. The
water molecules adsorb flat onto the surface at a Ca–O distance
of 2.47 Å and H–F distances of 2.13–2.18 Å. As suggested by
the DFT calculations, the Ca–Ca interatomic spacing of 3.85 Å
is too small for a water molecule to adsorb on each calcium ion,
and hence only 50% of the available adsorption sites are
covered by water molecules. We again calculated the adsorp-
tion energies with respect to isolated gaseous water molecules
to enable direct comparison with experimental techniques such
as temperature programmed desorption (TPD). We calculated
a hydration energy of 261.8 kJ mol21 at a partial coverage of
50%, which is in acceptable agreement with the DFT result of
253.4 kJ mol21. The discrepancy in the hydration energies is
due to the fact that the atomistic simulations predict a
completely flat mode of adsorption for the water molecules,

Table 2 Surface energies of the calcium fluoride surfaces

Surface energies/J m22

Surface Unhydrated Hydrated

{111} 0.52 0.40


{011} 0.82 0.90
Fig. 5 (a) Acute and (b) obtuse steps on the CaF2 {111} surface (Ca ~
{310} 1.56 0.67
black, F ~ pale grey).

J. Mater. Chem., 2003, 13, 93–101 97


Table 3 Adsorption energies of water and methanoic acid at the
surfaces

Adsorption energies/kJ mol21

Methanoic acid
Surface Water Methanoic acid in water

Planar {111} 238.5 256.3 296.7


Acute {111} 229.1 290.8 234.1
Obtuse {111} 250.8 279.5 221.2
{011} 233.4 2102.4 —
{310} 2250.7 2110.9 —

fashions. On the {011} surface, the lattice spacing is large


enough to allow full monolayer coverage with one methanoic
acid molecule per surface calcium. The methanoic acid
molecule adsorbs with both oxygen ions to two surface calcium
ions, bridging between them (Fig. 6). The doubly bonded Fig. 8 Plan view of the {111} surface with adsorbed methanoic acid
oxygen ion closely coordinates to the calcium ion at a distance molecule, showing the crystal as a lattice framework (Ca ~ black, F ~
of 2.2 Å while the oxygen atom of the hydroxyl group is at pale grey) and the methanoic acid molecule (space-filled O ~ black,
2.65 Å from the second calcium ion. The hydrogen of the C ~ pale grey, H ~ white).
hydroxyl group relaxes into the surface and coordinates to a
fluorine atom at 2.4 Å. The lattice spacing on the {310} surface to calcium ions further away in the next layer (3.93 Å).
is also large enough to accommodate full monolayer coverage Furthermore, the doubly bonded oxygen ions coordinate to
(Fig. 7). The doubly bonded oxygen ion is again bonded to a hydroxyl hydrogens of other adsorbed methanoic acid
surface calcium ion at 2.15 Å, while more loosely coordinated molecules (2.4 Å), while the hydrogen bonded to the carbon
atoms coordinates to surface fluoride ions (2.35 Å).
Due to the much smaller interatomic distance on the {111}
surface, only a 50% coverage of methanoic acid can be
accommodated, which is a reasonable coverage if we compare
it with experimental work by Mielczarski et al., who observed a
30% coverage of oleic acid, which is a carboxylic acid with a
long carbon chain instead of the hydrogen of methanoic acid.47
The molecules adsorb in a fairly flat configuration onto the
surface, bridging between two calcium ions, with both oxygen
ions coordinated to a calcium at 2.2 Å for the doubly bonded
oxygen ion and at 2.9 Å for the oxygen ion of the hydroxyl
group (Fig. 8). The hydrogen atom of the hydroxyl group
coordinates to two surface fluoride ions at 2.5 and 2.7 Å. When
adsorbed at the step edges, we see that the trend in adsorption
energies is reversed from the hydration pattern. More energy is
now released upon adsorption at the acute step edge than at the
obtuse edge, while both steps are calculated to be more
Fig. 6 Side view of the {011} surface, showing the crystal as a lattice favourable adsorption sites than the terraces of the planar
framework (Ca ~ black, F ~ pale grey) and the methanoic acid surface. The reason for the higher exothermicity at the acute
molecule (space-filled, O ~ black, C ~ pale grey, H ~ white) step edge is the fact that in addition to the same interactions
coordinated by its oxygens to two surface calcium ions.
between methanoic acid and the terrace atoms, as shown for
adsorption on the planar surface, the doubly bonded oxygen
atom also bonds to a low-coordinated calcium ion on the step
edge, hence bridging the gap between step and terrace, which
was not possible in the adsorption of water. The hydrogens also
interact with fluoride ions both on the edge and the terrace,
leading to a network of hydrogen-bonding between the surface
and adsorbate. It is these multiple interactions that cause the
methanoic acid adsorption at the steps to be more exothermic
than on the planar surfaces.

3.4 The effect of temperature: Molecular Dynamics simulations


of the (111) surface
As energy minimisation techniques do not take into account
temperature, in this section we employ Molecular Dynamics
simulations (MD) to explicitly investigate the effect of tempera-
ture on the interaction of water and methanoic acid at the
{111} plane at 300 K. We started the simulations with a
monolayer of water adsorbed at the surface. However, during
Fig. 7 Side view of the {310} surface, showing the crystal as a lattice
framework (Ca ~ black, F ~ pale grey) and the methanoic acid the simulation water molecules desorbed from the surface and
molecule (space-filled, O ~ black, C ~ pale grey, H ~ white) closely diffused through the gap between the two surfaces, forming
coordinated by its doubly bonded oxygen atom to a surface calcium ion a droplet (Fig. 9). From the mean square deviation, the
and hydrogen-bonding to a surface fluorine. diffusion coefficient of the water molecules is calculated to be

98 J. Mater. Chem., 2003, 13, 93–101


effect of temperature, included in calculations through the
employment of Molecular Dynamics simulations, is to exacer-
bate the difference in adsorption energies between water and
methanoic acid at the fluorite (111) surface. The MD simu-
lations also showed that although the methanoic acid diffuses
along the surface, it remains adsorbed, while competitive
interactions between the water molecules themselves outweigh
interactions between the solid surface and water molecules,
which as a result leave the surface.
Of course, in ‘‘real’’ mineral separation processes, water
and the organic flotation reagents co-exist and we have hence
extended our calculations to include both water and methanoic
acid in the simulations.

3.5 Co-adsorption of water and methanoic acid


The adsorption energies from the energy minimisations for
both water and methanoic acid onto the three surfaces are
collected in Table 3. The hydration energies for the different
calcium fluoride surfaces vary considerably due to the presence
Fig. 9 Water droplet formation between two {111} surfaces in a (or absence) of hydrogen-bonding to the surface fluoride ions,
Molecular Dynamics simulation at NVT and 300 K. in addition to the calcium–oxygen interactions. The adsorption
energies for methanoic acid onto the mineral surfaces are
1.2 6 1029 m2 s21, which is identical to that calculated for considerably larger than the hydration energies on the domi-
water molecules in a box of pure liquid water at 300 K under nant {011} and planar and stepped {111} surfaces due to the
NPT conditions.38 The hydration energy of 232.7 kJ mol21 at capability of the acid molecules to bridge, by their oxygen
300 K is less than the intermolecular interactions between atoms, between two or more surface calcium atoms and the
water molecules themselves, calculated to be 243.0 kJ mol21 close hydrogen-bonding to surface fluoride ions. These calcu-
at the same temperature in agreement with experiment lations would therefore suggest that it would be energetically
(243.4 kJ mol21).48 Hence, at this low water density, we preferential for the methanoic acid molecules to adsorb to these
may conclude that on energetic grounds the water molecules on surfaces, displacing the water molecules from the adsorption
the fluorite {111} plane prefer not to adsorb to the surface but sites, which was borne out by the MD simulations of water and
to cluster together. We next repeated the calculation with methanoic acid at the {111} surface at 300 K, where the water
methanoic acid as the adsorbate rather than water. This time, desorbed from the surface while the methanoic acid remained
however, the methanoic acid remained bound to the {111} adsorbed. However, in order to verify whether this assumption
surface rather than diffuse through the gap. Clustering of the based on separate calculations of the adsorbates is valid, we
repeated the calculations of methanoic acid adsorption at the
methanoic acid molecules also took place, but unlike the
planar and stepped {111} surfaces, but this time including a
water molecules, surface diffusion of the methanoic acid was
layer of water in the simulations, hence studying the compe-
followed by the formation of clusters at the surface, in a similar
titive adsorption of water and methanoic acid directly. The
adsorption pattern as was observed in the energy minimi-
adsorption energies were now calculated with respect to the
sation calculations above, by both oxygen atoms to surface
hydrated surface and a solvated methanoic acid molecule. In
calcium ions (Fig. 10). Similar clustering of methanoic acid was
order to model the co-adsorption of methanoic acid and water
observed experimentally by Iwasawa et al. to occur at terraces at the fluorite surfaces, we needed to derive interatomic
on the TiO2 (110) surface.49 The energy of interaction of potential parameters for the water–methanoic acid interac-
methanoic acid with the surface is calculated at 291 kJ mol21, tions. We fitted these potential parameters to the experimental
higher than in the energy minimisation simulations. Hence, the solvation energy of methanoic acid, in a series of Molecular
Dynamics simulations of a methanoic acid molecule in a box of
255 water molecules (Fig. 11). The final parameters thus
derived (Table 1) gave a solvation energy for methanoic acid
of 241.7 kJ mol21, compared to the experimental value of
247.4 kJ mol21.40
The data listed in Table 3 show that on the planar {111}
surface, the presence of water increases the adsorption energy
of methanoic acid, the reason for which becomes clear if
we compare the adsorption pattern of the methanoic acid
with that of water at the same surface sites. The methanoic acid
only replaces one adsorbed water molecule at the planar
surface and as the intermolecular interactions between the
water molecules themselves (43 kJ mol21) or with the
methanoic acid (40 kJ mol21) are very similar, the water
molecules have no preference for interacting with either the
methanoic acid or each other. The regular adsorption pattern
of the water on the surface is not disturbed by the presence of
the surfactant, but the adsorbate is stabilised by the forma-
tion of a network of hydrogen-bonded interactions to neigh-
bouring water molecules. However, at the stepped surface sites
the co-adsorption of water lowers the adsorption energies for
Fig. 10 Adsorption and clustering of methanoic acid molecules at the methanoic acid (Table 3), both processes becoming much less
{111} surface in a Molecular Dynamics simulation at NVT and 300 K. exothermic. Again, the reason is two-fold, based on both the

J. Mater. Chem., 2003, 13, 93–101 99


have presented calculations of the adsorption of water and
methanoic acid at calcium fluoride surfaces, using a combina-
tion of computational techniques. Accurate density functional
theory calculations were used to obtain the electronic struc-
ture of the {111} surface together with hydration modes
and energies; and atomistic simulation techniques to elucidate
the geometry and relative adsorption energies of water and
methanoic acid at a range of different surface sites. From our
simulations we can draw the following conclusions.
Electron density plots generated by DFT calculations of the
{111} surface show calcium fluoride to be a strongly ionic
crystal, with no discernible distortion in the surface layer with
respect to bulk layers, leading to minimal ionic relaxation of
the surface.
Associative adsorption of water is preferred at the {111}
surface, without significant energy barrier to reformation of
dissociated water molecules into molecular water. The hydra-
tion energy is dependent upon coverage and both the decrease
in hydration energy and lesser coordination to the surface upon
increasing coverage indicates repulsive interactions between the
adsorbed water molecules.
Fig. 11 Average configuration of methanoic acid molecule in a Modelling hydration of the {111} surface using atomistic
simulation cell of 255 water molecules during a Molecular Dynamics simulation techniques gives similar hydration energies and
simulation at NPT and 300 K (the apparently dissociated water configurations of the adsorbed water molecules to the DFT
molecules are, in fact, water molecules, but shown split up as an artefact
calculations, although there may be some overbonding of the
of the periodic boundary conditions).
H–F hydrogen-bonding in the interatomic potential approach.
Adsorption of methanoic acid up to full monolayer coverage
geometry of the surface sites and the relative adsorption
is possible on both {011} and {310} surfaces, but on the {111}
energies of the surfactant and the water molecules. Hydration
surface, due to the smaller calcium–calcium distance, only
of the stepped sites (29–51 kJ mol21) is energetically similar to
adsorption up to 50% is preferred. On this surface, the
the planar fluorite surface (average y38.5 kJ mol21). Binding
methanoic acid molecules adsorb by their oxygen atoms to two
of the methanoic acid to the steps is stronger than water
calcium atoms, forming a bridge between them. This mode of
(average y85 kJ mol21) and it therefore remains closely bound
adsorption is particularly favourable, which is also seen at the
to the step site even in the presence of water, as shown in
stepped sites on the {111} surface. On both the {011} and the
Fig. 12 for the obtuse step. However, its presence at the step
dominant {111} fluorite surfaces the energies of adsorption of
disturbs the regular pattern of water adsorption, at least in the
methanoic acid compared to water show that adsorption of
immediate vicinity of the step, leading to a smaller adsorption
methanoic acid is energetically more favourable and hence
energy. Thus, before the addition of water to the system we find
methanoic acid should compete effectively with water for
similar adsorption energies for the three different surface sites
adsorption at these surfaces. Molecular Dynamics simulations
on the {111} surface (56–91 kJ mol21), but once water has been
of the two adsorbates at the {111} surface show that the
introduced in the calculations, the adsorption energies show a
effect of temperature is to widen the gap in adsorption energies
much bigger variation with adsorption site (21–97 kJ mol21)
between methanoic acid and water. The latter adsorbate leaves
and even a reversal of the relative stabilities, indicating that we
the surface and forms a water droplet between the {111}
need to include solvent effects explicitly if we are to predict
planes.
realistic adsorption behaviour.
Simulations of methanoic acid adsorption in the presence of
water bear out the suggestion that methanoic acid competes
effectively with water as an adsorbate, as the methanoic acid
4 Conclusions remains adsorbed at the {111} terrace and steps forming close
interactions with the surrounding water molecules. However,
We have shown in this work that computational techniques are
these latter calculations have also shown that interactions
well placed to provide insight at the atomic level into the
between surfactant and water molecules can have a radical
interactions between substrate and adsorbate molecules. We
effect on adsorption behaviour, and it is therefore not suffi-
cient to calculate the interactions of surfactant molecules with
mineral surfaces in isolation, as the presence of solvent in the
calculations makes a significant contribution to the final
adsorption energies and relative stabilities of the surface sites.
The implication of these findings for the search for flotation
reagents is that we need to explicitly include solvent in the
calculations if we are to successfully predict the affinity of the
mineral for particular surfactants.

Acknowledgements
We acknowledge the Engineering and Physical Sciences
Fig. 12 Co-adsorption of water and methanoic acid at the obtuse step
Research Council, grant no. GR/N65172/01, and the Royal
on the {111} surface (fluorite as framework, methanoic acid space- Society, grant no. 22292, for funding and the Minerals and
filled, water as triangles; Ca ~ dark grey, F ~ pale grey, O ~ black, Materials Consortia for the provision of computer time on the
C ~ grey, H ~ white). Cray T3E.

100 J. Mater. Chem., 2003, 13, 93–101


References 27 N. H. de Leeuw and J. A. Purton, Phys. Rev. B, 2001, 63, 5417.
28 J. Goniakowski and M. J. Gillan, Surf. Sci., 1996, 350, 145.
1 E. S. Dana, in Manual of Mineralogy, John Wiley & Sons, New 29 M. Born and K. Huang, in Dynamical Theory of Crystal Lattices,
York, 1941. Oxford University Press, Oxford, 1954.
2 M. L. Free and J. D. Miller, Langmuir, 1997, 13, 4377. 30 G. W. Watson, E. T. Kelsey, N. H. de Leeuw, D. J. Harris and
3 K. H. Rao, J. M. Cases, P. de Donato and K. S. E. Forssberg, S. C. Parker, J. Chem. Soc. Faraday Trans., 1996, 92, 433.
J. Colloid Interface Sci., 1991, 145, 314. 31 N. H. de Leeuw, G. W. Watson and S. C. Parker, J. Phys. Chem.,
4 K. H. Rao, J. M. Cases and K. S. E. Forssberg, J. Colloid Interface 1995, 99, 17219.
Sci., 1991, 145, 330. 32 T. R. Forester and W. Smith, DL_POLY user manual CCLRC;
5 S. Parker, K. S. Songs, C. R. A. Catlow and A. M. Stoneham, Daresbury Laboratory, Daresbury, Warrington, UK, 1995.
J. Phys. C—Solid State Physics, 1981, 14, 4009. 33 L. Verlet, Phys. Rev., 1967, 195, 98.
6 J. H. Harding, Phys. Rev. B, 1985, 32, 6861. 34 S. Nosé, J. Chem. Phys., 1984, 81, 511.
7 D. G. Byrne, J. Corish and D. A. MacDonaill, J. Chem. Soc. 35 W. G. Hoover, Phys. Rev. A, 1985, 31, 1695.
Faraday Trans., 1999, 86, 1193. 36 C. R. A. Catlow, M. J. Norgett and A. Ross, J. Phys. C, 1977, 10,
8 H. Moller and H. E. L. Madsen, J. Crystal Growth, 1985, 71, 673. 1630.
9 R. Bennewitz, M. Reichling and E. Matthias, Surf. Sci., 1997, 387, 37 P. Osguthorpe, D. Osguthorpe and A. Hagler, Proteins; Structure,
69. Function and Genetics, 1988, 14, 31.
10 G. Jordan and W. Rammensee, Surf. Sci., 1997, 371, 371. 38 N. H. de Leeuw and S. C. Parker, Phys. Rev. B, 1998, 58, 13901.
11 S. Klump and H. Dabringhaus, Surf. Sci., 1998, 417, 323. 39 K. P. Schröder, J. Sauer, M. Leslie and C. R. A. Catlow, Chem.
12 G. Kresse and J. Hafner, Phys. Rev. B, 1993, 47, 5858. Phys. Lett., 1992, 188, 320.
13 G. Kresse and J. Hafner, Phys. Rev. B, 1994, 49, 14251. 40 D. R. Lide, in Handbook of Chemistry and Physics, CRC Press,
14 G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15. Boca Raton, USA, 81st edn., 2001.
15 G. Kresse and J. Furthmüller, Phys. Rev. B, 1996, 54, 11169. 41 W. A. Deer, R. A. Howie and J. Zussman, in An Introduction to the
16 R. Jones and O. Gunnarson, Rev. Mod. Phys., 1989, 61, 689. Rock Forming Minerals. Longman, Harlow, UK, 1992.
17 M. Payne, M. Teter, D. Allan, T. Arias and J. Joannopoulos, Rev. 42 C. A. Scamehorn, A. C. Hess and M. I. McCarthy, J. Chem. Phys.,
Mod. Phys., 1992, 64, 1045. 1993, 99, 2786.
18 M. Gillan, Contemp. Phys., 1997, 38, 115. 43 M. A. Johnson, E. V. Stefanovich, T. N. Truong, J. Gunster and
19 K. Refson, R. A. Wogelius, D. G. Fraser, M. C. Payne, M. H. Lee D. W. Goodman, J. Phys. Chem. B, 1999, 103, 3391.
and V. Milman, Phys. Rev. B, 1995, 52, 10823. 44 W. Langel and M. Parrinello, Phys. Rev. Lett., 1994, 73, 504.
20 P. J. D. Lindan, J. Muscat, S. Bates, N. M. Harrison and 45 J. Goniakowski and C. Noguera, Surf. Sci., 1995, 330, 337.
M. Gillan, Faraday Discuss., 1997, 106, 135. 46 N. H. de Leeuw, G. W. Watson and S. C. Parker, J. Chem. Soc.
21 P. J. D. Lindan, N. M. Harrison and M. J. Gillan, Phys. Rev. Lett., Faraday Trans., 1996, 92, 2081.
1998, 80, 762. 47 J. A. Mielczarski, E. Mielczarski and M. Cases, Langmuir, 1999,
22 S. P. Bates, G. Kresse and M. J. Gillan, Surf. Sci., 1998, 409, 336. 15, 500.
23 D. Vanderbilt, Phys. Rev. B, 1990, 41, 7892. 48 Z. Duan, N. Moller and J. H. Weare, Geochim. Cosmochim. Acta,
24 G. Kresse and J. Hafner, J. Phys.: Condens. Matter, 1994, 6, 8245. 1995, 59, 3273.
25 H. J. Monkhorst and J. D. Pack, Phys. Rev. B, 1976, 13, 5188. 49 Y. Iwasawa, H. Onishi, K. Fukui, S. Suzuki and T. Sasaki,
26 J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, Faraday Discuss., 1999, 114, 259.
M. R. Pederson, D. J. Singh and C. Fiolhas, Phys. Rev. B, 1992,
46, 6671.

J. Mater. Chem., 2003, 13, 93–101 101

You might also like