Você está na página 1de 4

J. Phys. Chem.

C 2007, 111, 18155-18158 18155

Modeling Aqueous Silica Chemistry in Alkali Media

Miguel J. Mora-Fonz, C. Richard A. Catlow, and Dewi W. Lewis*


Department of Chemistry, UniVersity College London, 20 Gordon Street, London WC1H 0AJ, United Kingdom
ReceiVed: September 6, 2007

A method for modeling the reactions of siliceous species is presented, which allows chemically accurate
deprotonation and dimerization energetics to be calculated, at conditions reflecting those of high pH
hydrothermal zeolite synthesis. The free energies of condensation reactions leading to silicate species up to
the linear tetramer are considered at room temperature and at 450 K.

The precise mechanism of zeolite formation is far from being ics of the deprotonation of both the monomer and the dimer,
completely understood, a consequence of the complex reaction together with those of the dimerization, trimerization, and
space and the difficulties in probing this chemistry under tetramerization reactions at 298 and 450 K. We also validate
reaction conditions. Notwithstanding such problems, the initial our method with the well-characterized water autoprotolysis
processes of nucleation and subsequent crystallization under reaction.
such hydrothermal conditions have been widely studied,1,2 and The calculations were performed using DMOL,3 version 2.2,15
key pre-nucleation species have been identified.3,4 We have also using a double numeric basis set plus polarization (DNP) and
recently demonstrated how computational methods can con- the BLYP functional. A description of the water solvent is
tribute to an understanding of the reactivity of silicate oligomers included via the COSMO16 approach, with initial gas-phase
in the formation of specific “zeolitic” building units, with the optimized structures being reoptimized in the presence of the
aim of establishing a more general picture of nucleation.5 solvation model. Molecular dynamics using DFT forces are also
However, key to progress in this area is establishing whether used to ensure initial models do not become trapped in local
computational models and methods are both accurate and also minima. The Gibbs free energy is calculated by combining the
reflect, as closely as possible, the conditions present during electronic energy (BLYP/DNP), zero-point energy, together with
synthetic and natural zeolite formation. the rotational, vibrational, and translational contributions to the
Silicic acid (Si(OH)4), the monomeric silicate molecule, is energy, using standard statistical mechanics methods at 298 K
the simplest but, perhaps, the most important species involved (the temperature of most of the experimental measurements)
in the formation of zeolite nucleation centers. Little zeolite and 450 K (characteristic of zeolite synthesis). Improvements
nucleation is noted under neutral conditions,2 and zeolite to this initial solvated model are made by the inclusion of
syntheses are typified by high pH (ca. 10-14). Two processes, sodium counterions in the charged clusters, together with the
in particular, are therefore vital in controlling the subsequent addition of some explicit (together with the continuum descrip-
chemistry: deprotonation of the monomer, Si(OH)4 + OH- f tion of COSMO) water molecules in the first coordination sphere
Si(OH)3O- + H2O (and further deprotonations of the resulting of all of the species considered, to represent more completely
anion), and dimerization by condensation of both neutral and the most important short-range interactions. Details of the
anionic species, for example, 2Si(OH)4 f (OH)3-Si-O-Si- number of explicit water molecules used for individual species
(OH)3 + H2O. are reported in Table 1, which gives the free energies calculated
Experimental measurements of monomer deprotonation and together with values derived from experimental pKa or equi-
dimerization energies have been reported and summarized by librium constants reported in the cited literature. Examples of
Iler6 and Brinker,7 with McCormick8 reviewing more recent the models used are shown in Figure 1.
measurements. Theoretical calculations have also been per-
The reliability of the methodology is established by consider-
formed using a variety of different models.9-14 The complexity
ing the autoprotolysis of water. It is clear that describing this
of silicate chemistry limits the experimental studies to only a
reaction either (most simplistically) by “gas-phase” models or
few species, specifically the monomer and dimer, as many
even with the inclusion of a COSMO solvation model drastically
different reactions occur once polymerization is initiated, the
overestimates the free energy change. However, the inclusion
individual speciation of which proves almost impossible.
of some explicit water molecules improves the description,
Similarly, computational methods have also been restricted by
giving the free energy within 4 kJ mol-1 of experiment. The
their relative high cost and also lack, in some instances, the
experimental value is obtained using what we consider the most
incorporation of descriptions of solvent and pH effects.
reliable current estimates of the entropies of the solution species
Here, we describe how a dielectric solvent model combined
given by Aue et al.,17 specifically that there is no difference in
with explicit hydration allows DFT methods to describe
entropy between a solvated water and a hydronium ion.
accurately the chemistry of small silicate oligomers, exemplified
here by Si(OH)4 and (OH)3-Si-O-Si-(OH)3, under conditions Inclusion of a full explicit hydration sphere does not appear
typical of zeolite formation. Specifically, we report the energet- necessary, as sufficient water is present to allow formation of
the stabilized complexes, with three and four water molecules
* Corresponding author. Tel.: +44 20 7679 4779. Fax: +44 20 7679 for the H+ and OH- ions, respectively. These complexes,
7463. E-mail: d.w.lewis@ucl.ac.uk. OH-(H2O)4 and H3O+(H2O)3, are found to be the most prevalent
10.1021/jp077153u CCC: $37.00 © 2007 American Chemical Society
Published on Web 11/15/2007
18156 J. Phys. Chem. C, Vol. 111, No. 49, 2007 Mora-Fonz et al.

TABLE 1: Free Energies of Reaction at 298 and 450 K (kJ mol-1) in the Gas Phase and COSMO Solvation (COSMO) as
Compared to Silicates Species up to the Tetramer with Experimental Results (expt)a
free energy (∆G)/kJ mol-1
298 K 450 K 298 K
COSMO
reaction gas COSMO exptb COSMO (Sefcik radii)
water autoprotolysis H3O+ + OH- f H2O + H2O -1013 -258 -261
H3O+(H2O)3 + OH-(H2O)4 f (H2O)4 + (H2O)5 -509 -97 -10120,21 -91
monomer I M + OH- f M- + H2O -226 -61 -64 -42
M + NaOH f MNa + H2O -90 -49 -44
M + NaOH(H2O)3 f MNa(H2O)3 + H2O -27 -29 -336; -368 -30 -29
monomer II M- + OH- f M2- + H2O 285 2 2 0
MNa + NaOH f MNa2 + H2O -64 -43 -43
MNa(H2O)3 + NaOH(H2O)3 f MNa2(H2O)6 + H2O -47 -17 -188; -1722; -236 -14 -16
dimer I D + OH- f D- + H2O -296 -91 -92 -64
D + NaOH f DNa + H2O -100 -79 -76
D + NaOH(H2O)3 f DNa(H2O)3 + H2O -27 -35 -388 -30
dimer II D- + OH- f D2- + H2O 119 -25 -23 -14
DNa + NaOH f DNa2 + H2O -73 -43 -40
DNa(H2O)3 + NaOH(H2O)3 f DNa2(H2O)6 + H2O -28 -25 -298 -24
dimerization M + M(H2O) f D(H2O) + H2O -10 -4 -78 -3
M + M f D + H2O -13 -2 0
M + MNa(H2O)3 f DNa(H2O)3 + H2O -12 -7 0
M + MNa f DNa + H2O -23 -32 -31
M + M- f D- + H2O -83 -33 -28
trimerization M + DNa f TrNa + H2O -42 -16 -8
M + DNa(H2O)3 f TrNa(H2O)3 + H2O -43 -10 -4
tetramerization M + TrNa f TNa + H2O -16 9 16
M + TrNa(H2O)3 f TNa(H2O)3 + H2O -18 -10 -6
a Monomers I and II are the first and second deprotonation reactions for the monomer, respectively, with similar notation used for the dimer. For

ease of comparison, those results for the most comprehensive model are in bold. Results are presented with standard COSMO atomic radii at both
298 and 450 K and for selected species at 298 K with the optimized radii given by Sefcik and Goddard13 (see text for further details). Monomer,
dimer, trimer, and tetramer are symbolized by M, D, Tr, and T, respectively. b The experimental free energy for the proton transfer between two
water molecules in solution is calculated from a thermodynamic cycle. The experimental enthalpy of formation of the species in the gas phase is
taken from Dewar et al.,20 and the change in entropy for a proton transfer is considered zero, after Aue et al.17 The free energies of hydration are
taken from Pearson.21

Figure 1. Models of the singly deprotonated monomer and dimer, and that of the doubly deprotonated dimer. The models depicted are the most
comprehensive considered here, with three explicit water molecules and a Na+ ion coordinated to the deprotonated oxygen, together with a dielectric
continuum model of the remaining solvent. We also considered the same silicate cluster but only surrounded by the continuum water model and
also simply as gas-phase species (see Table 1). The structures shown are the energy minima for the models shown, and dotted lines indicate
hydrogen bonds.

species in such solutions.18,19 Little change in either geometry ics as compared to experiment. It is only when both explicit
or energetics is found when further explicit water molecules solvation and the Na+ counterion are included in our model,
are added. making the system considered overall neutral, that the experi-
The first deprotonation energy (as pKa) of the monomer is mental deprotonation energies are described accurately (within
the most reliably determined thermodynamic property of silicate 4 kJ mol-1).
oligomers. Thus, this value serves as a key reference in While there is general agreement that deprotonation energies
evaluating the reliability of any theoretical method whose aim calculated with DFT-COSMO correlate well with experimental
is to investigate the self-assembly of silicate oligomers. From pKa and that it is possible to estimate pKa values by linear
our results, it is evident that the gas-phase and the COSMO extrapolation,9,22,23 DFT-COSMO alone, using the standard
models are again inadequate to model correctly the monomer’s atomic radii to compute the solvation cavity, is not considered
first deprotonation energy. Similarly, simply including the able to reliably compute absolute pKa values.9 Here, however,
counterion (Na+) together with COSMO results in poor energet- through the inclusion of some explicit solvation and counterions,
Modeling Aqueous Silica Chemistry in Alkali Media J. Phys. Chem. C, Vol. 111, No. 49, 2007 18157

good agreement with experiment is achieved. Thus, such an places these atoms “inside” the molecule and reduces the errors
approach, whereby some, but not necessarily all, of the inner in the cavity shape and size. Similar conclusions have been
solvation sphere of a deprotonated species is considered, may drawn before for silicates13 and for related systems, for example,
be more generally applicable, as discussed further below. concentrated aluminum hydroxide solutions.24,25
The presence of multiple species in solution makes accurate An alternative to the inclusion of explicit solvent, to improve
experimental determination of the pKa of the dimer significantly quantitative agreement with experiment, is to refit the radii
more difficult. Thus, values tend to be extrapolated from a used.13 However, such a fitting procedure rescales the radii of
composite value following a fitting procedure to determine the all atoms of the same element. Here, the most significant change
relative populations of the monomer and dimer.8 Again, there in atomic size is due to the deprotonation of the oxygen atoms
is good agreement (within 4 kJ mol-1) between our calculated in the silicate fragments. The semi-ionic nature of these species
deprotonation energies (once explicit hydration is considered) means that the charge is fairly well localized on this particular
and the available experimental results, giving both added oxygen. Hence, rescaling the radius of all oxygen atoms is
confidence in our method but also in the robustness of the unrealistic. Furthermore, the fitted radii are dependent on the
experimental fitting procedure, in this case. level of theory used,13 which will lead to some compromise in
Higher deprotonation energies have also been determined for transferability. To illustrate that our method of inclusion of
the monomer together with the first and second deprotonation explicit water, which in essence ensures that changes in atomic
energies for the dimer.24 However, these rely increasingly on size do not result in a poorer representation of the cavity, is
fitting procedures, where the concentrations of the more strongly both transferable and, indeed, gives better agreement with
deprotonated species and of the dimer species are very small experiment, we present in Table 1 the results obtained without
as compared to the dominant singly and doubly deprotonated explicit water but with the Sefcik and Goddard13 COSMO radii.
monomeric species. Hence, these values are likely to be less These authors used B3LYP 6-31G**+//6-31G** for their
reliable. Nevertheless, it is clear that the first and second calculations, and they note that the fitted radii are therefore
deprotonation energies of the dimer are well described with our specific to this level of theory. Hence, our deprotonation energies
method (Table 1), giving us increased confidence in modeling using their radii but with the BLYP/DNP/COSMO method do
larger silicate species. not exactly reproduce the experimental values. On the contrary,
when the Sefcik and Goddard radii are used together with
The dimerization reaction is the fundamental reaction of silica
explicit water, the results are almost comparable to our original
chemistry. Again, the range of conditions (acidic, neutral, basic)
results (and hence experiment). Thus, our hypothesis that the
and the difficulties in attempting to restrict further polymeri-
explicit water provides an effective shielding of the ionic species
zation occurring make it a challenging reaction to study
is supported.
experimentally. The experimental estimate of the ∆G is = -7
We acknowledge that the level of theory considered here
kJ mol-1.8 In addition, there are a number of computed values:
should not, perhaps, be expected to give as good an agreement
using DFT/TNP, Catlow et al.12 reported a gas-phase dimer-
as found. Nevertheless, the method performs consistently for
ization free energy of -9.2 kJ mol-1, while Tosell,11 using MP2/
all of the species and conditions considered and at a relative
6-311+G(2d,p), reported a value of +24.7 kJ mol-1 in the gas
low computational cost. Indeed, the inclusion of the combination
phase but +8.8 kJ mol-1 when a G2/COSMO model is applied.
of explicit and COSMO solvation appears to provide a good
Most recently, Trinh et al. reported a value of +9 kJ mol-1 for
compromise between expense and accuracy. Note that, at this
the neutral dimerization and -28 kJ mol-1 for the reaction of
level of theory, the largest cluster presented here (the hydrated
a monomer with a deprotonated monomer, both using B3LYP/
tetramer) requires 25 000 min on a 2.8 GHz Xeon processor,
6-31+G(d,p) with COSMO solvation.23 However, none of these
with 2 GB memory. Furthermore, little change is noted, for
studies consider explicit hydration.
example, when the number of explicit waters is increased further
Our calculations reproduce well the experimental data for the here, nor when Car Parinello quantum molecular dynamics were
dimerization reaction, both under neutral and under basic used (at considerable higher cost) to consider aluminate spe-
conditions. Note, that these calculations correctly reflect how cies.25 Hence, the method, which allows us to consider condi-
an increase in pH makes polymerization more favorable. tions that are chemically relevant to zeolite synthesis, can, we
Moreover, we find the reaction of two deprotonated monomers believe, be applied with confidence to further silicate species.
(with cations and explicit water) to be slightly less favorable We show, for example, in Table 1 the results for the reactions
(-3.4 kJ mol-1), again reflecting experimental observation that discussed above at 450 K, a temperature akin to typical zeolite
very high pH does not necessarily result in increased polym- synthesis, and also for the trimerization and tetramerization
erization. Thus, we believe that the combination of the DFT reactions. Of note here is that there is little variation in the free
method used, together with a suitable explicit representation of energy of adding a monomer to a hydrated monomer, dimer,
the inner solvation sphere of the key species and a continuum or trimer at 298 K. However, the formation of such small
representation of the remainder of the solvent, allows the fragments through monomer addition is less favorable at
aqueous chemistry of silicate oligomers to be modeled reliably. elevated temperatures, suggesting that lower temperatures are
The effect of including explicit water molecules can be required to initiate formation of larger silicates. Indeed, we note
considered two-fold. First, they shield the highly charged regions that room-temperature aging of synthesis gels is often critical
from the continuum, allowing electrons to redistribute more to the successful formation of zeolite crystals.
“naturally”. Second, they result in an improvement in the Further studies using this method will allow us to approach
description of the shape of the cavity in the dielectric continuum a more comprehensive description of the key reactions that occur
that is constructed by the COSMO methodology. COSMO uses during the pre-nucleation phase of zeolites and other silicates.
atomic radii to construct the cavity,16 and the large difference
in atomic size between neutral atoms and these charged and Acknowledgment. We acknowledge Dr. Jörg A. Becker and
ionic species is significant and results in a poorer description Dr. Jürgen Woenckhaus for stimulating discussion regarding
of the surface of the molecule. Thus, addition of explicit water this work. M.J.M.-F. is greatly indebted to CONACYT Mexico
near such atoms, particularly bare O- and Na+, effectively and UCL for financial support. These calculations were
18158 J. Phys. Chem. C, Vol. 111, No. 49, 2007 Mora-Fonz et al.

performed on a computational resource provided by the EPSRC (13) Sefcik, J.; Goddard, W. A. Geochim. Cosmochim. Acta 2001, 65,
to D.W.L. through grant GR/S06233/01. 4435.
(14) Trinh, T. T.; Jansen, A. P. J.; van Santen, R. A. J. Phys. Chem. B
2006, 110, 23099.
References and Notes
(15) Delley, B. J. Chem. Phys. 2000, 113, 7756.
(1) Cundy, S. C.; Cox, A. P. Chem. ReV. 2003, 103, 663. (16) Baldridge, K.; Klamt, A. J. Chem. Phys. 1997, 106, 6622.
(2) Cundy, C. S.; Cox, P. A. Microporous Mesoporous Mater. 2005, (17) Aue, D. H.; Webb, H. M.; Bowers, M. T. J. Am. Chem. Soc. 1976,
82, 1. 98, 318.
(3) Kinrade, S. D.; Knight, C. T. G.; Pole, D. L.; Syvitski, R. T. Inorg. (18) Tuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M. J. Chem.
Chem. 1998, 37, 4272. Phys. 1995, 103, 150.
(4) Kirschhock, C. E. A.; Kremer, S. P. B.; Grobet, P. J.; Jacobs, P.
A.; Martens, J. A. J. Phys. Chem. B 2002, 106, 4897. (19) Tuckerman, M. E.; Marx, D.; Parrinello, M. Nature 2002, 417, 925.
(5) Mora-Fonz, M. J.; Catlow, C. R. A.; Lewis, D. W. Angew. Chem., (20) Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. F.; Stewart, J. J. P. J.
Int. Ed. 2005, 44, 3082. Am. Chem. Soc. 1985, 107, 3902.
(6) Iler, R. K. The Chemistry of Silica: Solubility, Polymerization, (21) Pearson, R. G. J. Am. Chem. Soc. 1986, 108, 6109.
Colloid and Surface Properties, and Biochemistry; Wiley: New York, 1979. (22) Tossell, J. A.; Sahai, N. Geochim. Cosmochim. Acta 2000, 64, 4097.
(7) Brinker, C. J.; Scherer, G. W. Sol-Gel Science: The Physics and
Chemistry of Sol-Gel Processing; Harcourt Brace Jovanovich: Boston, 1990. (23) Mora-Fonz, M. J.; Catlow, C. R. A.; Lewis, D. W. Stud. Surf. Sci.
(8) Sefcik, J.; McCormick, A. V. AIChE J. 1997, 43, 2773. Catal. 2005, 158, 295.
(9) Schuurmann, G. J. Chem. Phys. 1998, 109, 9523. (24) Caullet, P.; Guth, J. L. Observed and Calculated Silicate and
(10) Pereira, J. C. G.; Catlow, C. R. A.; Price, G. D. Chem. Commun. Aluminosilicate Oligomer Concentrations in Alkaline Aqueous Solutions.
1998, 1387. In Zeolite Synthesis; Occelli, M. L., Robson, H. E., Eds.; American Chemical
(11) Tossell, J. A. Geochim. Cosmochim. Acta 2005, 69, 283. Society: Washington, DC, 1989.
(12) Catlow, C. R. A.; Coombes, D. S.; Lewis, D. W.; Pereira, J. C. G. (25) Sillanpää, A. J.; Päivärinta, J. T.; Hotokka, M. J.; Rosenholm, J.
Chem. Mater. 1998, 10, 3249. B.; Laasonen, K. E. J. Phys. Chem. A 2001, 105, 10111.

Você também pode gostar