You are on page 1of 12

Journal of Mineralogical and dissolution

Silica Petrological Sciences,


catalyzed byVolume
NaOH 107, page 87 ─ 98, 2012 87

Silica dissolution catalyzed by NaOH: Reaction kinetics


and energy barriers simulated by
quantum mechanical strategies

Osamu Tamada*, Gerald V. Gibbs**, Monte B. Boisen Jr.***


and J. Donald Rimstidt**
*
Department of Bio-Environmental Sciences, Kyotogakuen University,
Sogabecho-Kameoka, Kyoto 621-8555, Japan
**
Department of Geological Sciences, Virginia Polytechnic Institute and
State University, Blacksburg, VA 24061, USA
***
Department of Mathematics, University of Idaho, Moscow, ID 83844-1103, USA

Quartz dissolves in Na+ solutions much faster than in pure water though the experimental activation energies are
essentially the same. To elucidate the mechanism, silica dissolution was simulated using Gaussian 03 software
at B3LYP/3-21G* initially and the B3LYP/6-311G(2d,p) levels, applied to a modeled silica Si4O6(OH)4 in the
presence of an NaOH molecule and up to five water molecules. Each water molecule was added one by one to
the system and approached the surface silicon in successive geometry optimizations. The atomic positions of
the sodium hydroxide, the water molecules and the surface SiO3(OH) moiety were varied to mimic the surface
reaction whereas the positions of the remaining Si3O3(OH)3 atoms were frozen to represent the bulk structure
throughout the geometry optimization.
   In the first step a single H2O molecule was added to the system [Si4O6(OH)4 + NaOH]. None of the Si-O bond-
ed interactions were ruptured by the intrusion of the water molecule but the surface Si was stabilized in the Q3Si site
(connected to three Si-O-Si bridges), being coordinated by five oxygen atoms. The energy barrier was 63 kJ/mol.
   In each of the second and third steps one more water molecule was introduced to the system. One Si-O
bonded interaction of the Si-O-Si bridges was ruptured to make Q2Si and Q1Si sites in the second and third
steps, respectively, and the energy barriers were low (22-29 kJ/mol).
In the fourth and fifth steps, the added water molecules were prevented by the sodium ion from reaching the last
Si-O-Si bridge, leaving the Si in the Q1Si site.
   The basis set was raised to the 6-311G(2d,p) level and applied to the 63 kJ/mol barrier found in the first
step to determine the maximum energy barrier in the series of reactions. The barrier increased to 88 kJ/mol (82
kJ/mol in enthalpy), which is still in the range of experimental activation energies of 46-96 kJ/mol.
In summary sodium works to stabilize the surface silicon in penta-coordination with the energy barrier, 82 kJ/
mol, resulting in longer Si-O distances and the weakening of the bonded interactions. This makes the Si-O rup-
ture easier and faster, which gives insight as to how the presence of alkali enhances silica dissolution.

Keywords: Sodium hydroxide, Gaussian 03, Activation energy, Geometry optimization, Alkali enhancement

INTRODUCTION bonded interaction is key to interpreting many important


geological and geochemical processes. For example, rock
Understanding the interaction of water with the Si-O strength is affected by hydrolytic weakening (e.g., Tullis
and Yund, 1980, 1989). The full range of water-rock in-
doi:10.2465/jmps.110909 teractions found in magma generation and evolution
O. Tamada, tamada@t02.mbox.media.kyoto-u.ac.jp Corresponding (Burnham 1997), metamorphism (Spear, 1993), wall rock
author
G.V. Gibbs, ggibbs@vt.edu
alteration (Reed, 1997) and chemical weathering (White
M.B. Boisen, boisen@uidaho.edu and Brantley, 1995) involve the reaction of water with
J.D. Rimstidt, jdr02@vt.edu Si-O bonded interactions. In the waters in the natural en-
88 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

vironment, alkali ions are commonly present and play an ment, the activation energies in the presence of alkali cat-
important role in water-rock interactions. It is, therefore, ions (71.2 kJ/mol for NaCl and 71.5 kJ/mol for KCl) are
of great value to investigate the interaction of Si-O bonds reported by them to be the same as that in the deionized
and water in the presence of alkali ions. solution (71.3 kJ/mol). Thus, the activation energies so far
Numerous experimental studies of the kinetics of sil- reported are in the same range as for silica dissolution in
ica dissolution reactions, reviewed by Dove and Rimstidt deionized water. The reason for the alkali enhancement is
(1994), provide base-line information about the nature of still unclear.
the silica- water interaction. In these studies the activation On the other hand, theoretical tools of computational
energies for quartz dissolution were estimated to be chemistry, such as molecular dynamics and molecular or-
roughly 75 kJ/mol (18 kcal/mol) although there is a large bital geometry optimization strategies, enabled the model-
scatter (46-96 kJ/mol) around this value (Table 1). It is ing of surface reactions at the atomic level (Lasaga, 1995;
well known that silica dissolution is endothermic reaction Felipe et al., 2001). In an earlier study, Lasaga and Gibbs
(Iler, 1979) and that the presence of alkali cations like Na (1990) pointed out that the reaction, H2O + ≡Si-O-Si≡ →
and K enhances the dissolution rate. For example, Dove 2 ≡Si-O-H, is an elementary step in the hydrolysis pro-
and Crerar (1990) determined the quartz dissolution rates cess occurring at the surface of silicate. Subsequent stud-
at near-neutral pH in the solutions with as little as 0.05 ies have been carried out in an attempt to clarify the silica
molal NaCl or KCl concentrations at the hydrothermal dissolution kinetics (Xiao and Lasaga, 1994, 1996; Walsh
conditions of 100 °C to 300 °C, and found that the rate is et al., 2000; Criscenti et al., 2006; Nangia and Garrison,
enhanced by a factor of 33. In spite of the alkali enhance- 2008, 2009).
Several studies have simulated the hydrolysis of sili-
Table 1. Experimental and calculated activation energies (Ea, kJ/
ca surfaces using molecular dynamics strategies (Cheng
mol) for silica dissolution et al., 2002; Du et al., 2003; Rignanese et al., 2004; Ma et
al., 2005). These studies indicate that the attack of water
at the defect sites on a silica surface lowers the energy
barrier and that the reactions are catalyzed by a coopera-
tive transfer of hydrogen atoms contributed by the water
molecules.
Pelmenschikov et al. (2000) considered the resis-
tance of the bulk structure to the activated surface and
found that the activation energy decreases as the “con-
nectedness” of the target Si decreases from Q4Si (con-
nected via four bridging oxygen atoms to four adjacent Si
atoms) to Q2Si (connected via two bridging oxygen atoms
to two adjacent Si atom). They concluded, in a later study
(Pelmenschikov et al., 2001), that the experimental acti-
vation energy correlates with the cleavage of the last Si-
O-Si bonded interaction in Q1Si. Cypryk and Apeloig
(2002) reported that the protonation of the siloxane oxy-
gen, assisted by the cooperative hydrogen transfer, was
the most important factor in reducing the activation ener-
gy barrier. Recently Wallace et al. (2010) investigated the
ability of M(II) aqua ions (Mg2+, Ca2+) to promote the hy-
drolysis of Si-O bonded interactions.
In these computational efforts we find some features
such as the calculated activation energies are generally
higher by roughly 40-80 kJ/mol than the experimental
ones as shown in Table 1. It is also suggested that they are
reduced by the “connectedness” (Pelmenschikov et al.,
2000), that the addition of water molecules to the dissolu-
tion system reduces the activation energy (Cypryk and
Note; (Qn → Qm ) indicates the rupture of Si-O-Si bridges con- Apeloig, 2002) and that the high and low pH values en-
nected to Si atom from n bridges to m bridges. hance the dissolution rate as indicated in the study of the
Silica dissolution catalyzed by NaOH 89

protonated, neutral and deprotonated species (Nangia and tion kinetics in the presence of sodium hydroxide, the
Garrison, 2008, 2009). second goal is to address how the calculated activation
The Na enhancement of the experimental silica dis- energies relate to the experimental ones for silica dissolu-
solution rate is markedly observed in solutions from neu- tion with sodium hydroxide.
tral to high pH (Dove, 1995). The enhanced rate curve ex-
tends smoothly from high to very high pH, which COMPUTATION AND RESULTS
suggests that Na enhances dissolution even in solutions
with very high pH. Consequently, we use the sodium hy- Gaussian 03 (Frisch et al., 2003) and Xtaldraw (Bartel-
droxide molecule for modeling the system of silica in ba- mehs et al., 1993) were used for the ab initio calculations
sic solution and for simplifying the system. and crystal structure drawings, respectively.
Sverjensky (2005) used a theoretical framework to At the silica-water interface, the silica surface is hy-
predict surface charges on oxides in 1:1 (M+;L−) electro- droxylated with the major part of the surface consisting of
lytes salt solutions and found that on high dielectric con- silanol in which the surface silicon is in the Q3Si site
stant solids, such as rutile, the alkali cations predominant- (connected to three Si-O-Si bridges and a silanol) (Iler,
ly adsorb close to the surface as inner-sphere complexes 1979; Parks, 1984). On this basis, the silica was modeled
but on most other oxides, including silica, they predomi- as a hydrogen clad, Si4O6(OH)4 molecule which was ge-
nantly adsorb further away as outer-sphere complexes. ometry optimized at the B3LYP/6-311G(2d,p) level, as-
This means that most adsorbed sodium ions on a silica suming S4 point symmetry. The resulting configuration is
surface are present as outer-sphere complexes. However, composed of four silicate tetrahedra corner sharing each
Sverjensky’s model implies that even though the concen- others. This molecule is used to model the silica surface
tration of adsorbed alkali cations present as inner-sphere and the matrix by considering it in two parts, one variable
complexes is small there is an equilibrium between inner- surface SiO3(OH) moiety that is flexible so that it can re-
and outer-sphere cations so that some inner-sphere com- spond to the water attack and three frozen inner tetrahedra
plexes are present on silica surfaces. Wallace et al. (2010) Si3O3(OH)3 matrix used to model the bulk structure.
investigated the effect of both inner- and outer-sphere
complexes of Mg2+ and Ca2+ on the hydrolysis of Si-O First step
bonded interactions and found that the height of the hy-
drolysis barrier is significantly lower for the inner-sphere The oxygen atoms of a water molecule and a sodium hy-
site. They concluded that, despite their documented abun- droxide were initially placed 4.3 Å and 3.5 Å respectively
dance at the silica-water interface, outer-sphere surface from the target surface silicon Sit atom as shown in Figure
complexes are not likely participants in the Si-O bonded 1. The atomic positions of each of the atoms in the
dissociation mechanism. Taking these points into account, SitO3(OH) moiety, the sodium hydroxide and the water
we designed a simulation model, which considers the ef- molecule were free to move during the geometry optimi-
fect of sodium ions adsorbed on silica surfaces as inner- zation with the constraint that the Sit-Ow distance r (Ow:
sphere complexes because they appear to be the species the water’s oxygen) was fixed at 4.3 Å. As indicated
most likely responsible for the observed increase in disso- above, the remaining Si3O3(OH)3 atoms associated with
lution rates. the interior of the silica model were frozen at their initial-
In this study, we examine the attack of multiple wa- ly optimized positions. The geometry optimization of the
ter molecules on a neutral silica surface in the presence of system was completed at the B3LYP/3-21G* level. After
an NaOH molecule. In this dissolution simulation, the the optimization, the Sit-Ow distance r was decreased by
strategy is to have each water molecule approach the sili- the value Δr, and as before, the positions of the atoms in
ca model Si4O6(OH)4, step by step producing the opti- the SitO3(OH) moiety, sodium hydroxide and the water
mized geometry at each step, thereby moving along the molecule were again relaxed during the energy optimiza-
energy valley. tion. This strategy was continued with a progressive
The objective of this simulation is to find what are shortening of the Sit-Ow distance from 4.3 to 1.55 Å. A
likely to be the intermediate structures and the values of graph of the total energies was obtained along the poten-
the intermediate energies in a series of reactions. In this tial surface valley of the Sit-Ow pathway as displayed by
way, we are able to obtain insight into the dissolution pro- the curve in Figure 2.
cess that is not possible to observe in experiments, and, The total energy increased from a distance of 3.4 to
therefore, to promote our understanding of silica dissolu- 2.3 Å with the slope of the energy curve progressively in-
tion catalyzed by sodium hydroxide. creasing with decreasing distance. From this energy curve
Besides our primary goal of clarifying the dissolu- and the geometries, estimates of the adsorption state and
90 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

the total energy decreased until a distance of 1.8 Å was


obtained. The product geometry was also optimized relax-
ing Sit-Ow distance as shown in Figure 3 together with the
geometries of adsorption and transition states. Note that
the five-fold coordinated silicon atom is sitting at the sta-
ble Q3Si site in the minimum energy configuration. The
sodium atom is connected to one of the three bridging ox-
ygen atoms, an oxygen originated from the NaOH mole-
cule and the water’s oxygen. As a result, the sodium atom
weakens the Si-O bonded interactions around the surface
Si and the Si-O distances are elongated from ~ 1.65 Å to
~ 1.75 Å, which are suitable bond lengths for a five-fold
coordinated silicon. In previous works, there has been no
report that the sodium stabilizes the silicon in the five-
fold coordination in the neutral surface condition.
Figure 1. The initial geometry for the first step. The modeled silica In order to highlight the role of sodium hydroxide in
Si4O6(OH)4, an NaOH molecule and a water molecule are shown
silica dissolution the same calculations were performed
in theTamada
figure. The
et al.bold line
Fig.separates
1 the surface part of
SitO3(OH) moiety, H2O and NaOH molecules from the matrix setting the model without sodium hydroxide (the model is
part of Si3O3(OH)3. The atoms in the surface portion are all acti- the same as above but the NaOH molecule is replaced by
vated and those in the matrix are all frozen throughout the opti- a water molecule HOH, in which each hydrogen is placed
mizations. about 1 Å from the water oxygen). A graph of the total
energies and the geometries of adsorption and transition
-2216. + states, and product are shown in Figures 4 and 5, respec-
‐0.495 tively. In this case the activation energy from the adsorp-
Transition
tion state to the transition one is fairly high (112 kJ/mol)
‐0.500
and one of three Si-O-Si bonded interactions is ruptured
to make Q2Si site as shown in the product. This illustrates
e)
Energy((Hartree

‐0.505
a clear difference from the process with sodium hydroxide
‐0.510
0 510 63 kJ/mol where the Si atom is stabilized in a five-fold coordination
and in Q3Si site. The silicon atom is never stabilized in
‐0.515 higher coordinations than four without sodium hydroxide
Product and this is apparently the role that sodium plays in silica-
‐0.520
water interactions.
‐0.525
Adsorption Second step
‐0.530
0 530
1.5 2.0 2.5 3.0 3.5 4.0 4.5 A second water molecule was added to the product geom-
Sit-Ow distance (Å) etry obtained in the first step and the same strategy as in
Figure 2. A graph of the total energies calculated at the B3LYP/3- the first step was applied. That is, the atomic positions of
Tamada et al.     Fig. 2
21G* level versus the distance r of the water oxygen Ow to the surface SitO3(OH) moiety, sodium hydroxide and two wa-
target silicon atom Sit in the first step. At each point the geometry ter molecules were free to move during the geometry op-
was optimized under the constraint of a fixed r distance. The ad-
timization. The distance between the target Si and the ox-
sorption state, transition state and product were optimized with
the distance constraint removed, starting from the geometries op- ygen Ow of the added water molecule were fixed and the
timized under the constraint. geometry was optimized. In this way the optimizations
were performed step by step with each step decreasing the
Sit-Ow distance along the energy valley of Sit-Ow pathway.
the transition state were found. The actual modeled ad- The energy curve is shown in Figure 6. The energy gener-
sorption and transition states were found by taking the ge- ally decreases with decreasing distance from 4.0 Å to 2.0
ometries of the estimated states and optimizing these re- Å (although there is a hump around 3 Å), goes up to a
laxing the Sit-Ow distance. The resulting values are plotted sharp peak at 1.8 Å and falls down to a minimum at 1.7 Å.
in Figure 2. The energy difference between these states From the geometries of these constrained optimizations
was 63 kJ/mol. As the Sit-Ow distance further decreased, the adsorption state, reactant, transition state and product
Silica dissolution catalyzed by NaOH 91

Adsorption Transition state Adsorption Transition state

H 2O Product H2O
H2O
H 2O
Product

Ow

Sit in Q3Si
Figure 5.
Tamada et The
al. geometries
Fig. 5 of the adsorption state (r = 3.6 Å), transi-
Ow tion state (r = 1.9 Å) and product (r = 1.7 Å) in the reactions
where the NaOH molecule was replaced by a water molecule
H2O. The product is characterized by the rupture of one of three
N
Na Si-Obr bonded interactions to make a Q2Si site keeping the four-
Figure 3. The geometries of the adsorption state (r = 3.4 Å), transi- fold coordination of Si.
Tamada et al.     Fig. 3
tion state (r = 2.3 Å) and product (r = 1.8 Å) in the first step. The
product is characterized by the five-fold coordinated silicon -2292. +
atom. ‐0.525

-2055. + ‐0.530 Transition


ee)

‐0.660
T
Transition
iti
gy(Hartre

‐0.670 ‐0.535
Hartree))

112 kJ/mol
Energ

‐0.680
Adsorption
‐0.540
Energy(H

‐0.690
22 kJ/mol

‐0.545
‐0.700
0.700 Reactant
E

Product
‐0.710
‐0.550
Product
Adsorption
1.5 2.0 2.5 3.0 3.5 4.0 4.5
‐0.720
1.5 2.0 2.5 3.0 3.5 4.0 4.5
Sit-Ow distance (Å)
Sit-Ow distance (Å) Figure 6. A graph of the total energies calculated at the B3LYP/3-
21G* level versus the distance r of the water oxygen Ow to the
Figure 4. A graph of the total energies for the system Si4O6(OH)4 +
2H2O without sodium hydroxide, calculated at the B3LYP/3- Tamada et al.     Fig. 6
target silicon atom Sit in the second step. At each point the geom-
Tamada
etry was optimized under the constraint of a fixed r distance. The
21G* level versus the et al. Fig.
distance r of 4the water oxygen Ow to the
adsorption state (r = 3.5 Å), reactant (r = 2.0 Å), transition state (r
target silicon atom Sit. At each point the geometry was optimized
= 1.8 Å) and product (r = 1.7 Å) were optimized without the dis-
under the constraint of a fixed r distance. The adsorption state,
tance constraint, starting from the geometries optimized under
transition state and product were optimized with the distance
the constraint.
constraint removed, starting from the geometries optimized un-
der the constraint.

energies. It is reasonable that this low activation energy


were successfully obtained at the distances of 3.5, 2.0, 1.8 results from the weakening of the Si-O bonded interac-
and 1.7 Å, respectively. The energy difference between tion associated with the five-fold coordination of silicon.
the reactant at 2.0 Å and the transition state at 1.8 Å is 22 The geometries of these states are shown in Figure 7. The
kJ/mol, a value far lower than the experimental activation added water molecule approaches the target silicon and
92 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

Adsorption Reactant -2368. +


0 545
‐0.545

gy(Hartrree)
H 2O ‐0.550
H 2O
29 kJ/mol
‐0.555 22 kJ/mol

Energ
Transition Fi t max.
First
‐0.560

‐0.565
0 565

Transition state Product TTamada et al. 


d l
‐0.570
Fig. 7 Reactant
Adsorption
Product
‐0.575
1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

Sit-Ow distance (Å)


Ow
Ow
Sit in Q2Si Figure 8. A graph of the total energies calculated at B3LYP/3-21G*
level versus the distance r of the water oxygen Ow to the target
silicon atom SiTamada et al.       Fig. 8
t in the third step. At each point the geometry was

Na optimized under the constraint of a fixed r distance. The adsorp-


tion state (r = 3.5 Å), first maximum (r = 2.7 Å), reactant (r = 2.0
Figure 7. The geometries of the adsorption state (r = 3.5 Å), reac- Å), transition state (r = 1.74 Å) and product (r = 1.72 Å) were
tant (r = 2.0 Å), transition state (r = 1.8 Å) and product (r = 1.7 Å) optimized without the distance constraint, starting from the ge-
in the second step. This step is characterized by the rupture of ometries optimized under the constraint.
one of the three Si-O-Si bridges, making the Q2Si site.
energy difference between the adsorption state at 3.5 Å
and the first maximum at 2.7 Å is 29 kJ/mol and that be-
makes a bonded interaction with the silicon momentarily tween the reactant at 2.0 Å and the transition state at 1.74
to form a six-fold coordinated silicon intermediate. It at- Å is 22 kJ/mol. Both of them are far below the experi-
tacks one of the three bridging oxygen atoms, transferring mental activation energies. It is clear that these low acti-
a hydrogen to it in the transition state. As a result, the vation energies result from the weakening of Si-O bonded
bridging oxygen moves from the Si atom, which consti- interaction associated with the five-fold coordination of
tutes the rupture of Si-Obr bonded interaction, leading to silicon. The geometries of these states are shown in Fig-
the formation of Q2Si site. At the end of this step, the so- ure 9. After the added water molecule has passed through
dium atom has caused a weakening of the Si-O bonded the distance of the first maximum as it approaches the tar-
interactions, connected to the three oxygen atoms around get silicon atom, it attacks one of the two Si-O-Si bridg-
the target silicon resulting in the five-fold coordinated sil- ing oxygen atoms yielding a hydrogen to that oxygen just
icon atom as shown in the product geometry. before the transition state as shown in the figure at r = 1.8
Å, at which time the target silicon is instantaneously coor-
Third step dinated with six oxygen atoms. As a result the bridging
oxygen moves from the Si atom rupturing the Si-O bond-
A third water molecule was added to the product geome- ed interaction, which leads to the formation of a Q1Si site.
try obtained in the second step and the same strategy as in After this has occurred, the target silicon atom is only
the previous steps was performed. The energy curve is connected with the inner matrix by the last Si-O-Si bridg-
shown in Figure 8. As the Si-Ow distance goes from 4.7 Å ing oxygen and is still coordinated by five oxygen atoms
to 3.5 Å , the energy decreases. The energy increases as as shown in the product geometry.
the distance decreases from 3.5 Å to around 2.7 Å. Then,
it decreases to a minimum at 2.0 Å, again increases to a Fourth and fifth steps
sharp peak at 1.74 Å and then drops down to a minimum
at 1.73 Å. From the geometries of these constrained opti- In each of the fourth and fifth steps one additional water
mizations the adsorption state, the first maximum (com- molecule was added. But these additional water molecules
putationally a transition state), the reactant, the transition were prevented by the sodium atom from reaching the last
state and the product were successfully obtained at the Si-O-Si bridge. It is possible that, after the sodium atom
distances 3.5, 2.7, 2.0, 1.74 and 1.72 Å, respectively. The is saturated with water molecules, the next one might suc-
Silica dissolution catalyzed by NaOH 93

Adsorption First maximum Reactant

H 2O
H 2O

H 2O

Before Si‐O rupture at  1.80 Å Transition state Product

Figure 9. The geometries of the ad-


sorption state (r = 3.5 Å), first max-
imum (r = 2.7 Å), reactant (r = 2.0
Å), the point just before Si-O rup-
ture (r = 1.8 Å), transition state (r =
1.74 Å) and product (r = 1.72 Å) in
H 2O Ow Sit in Q1Si Ow the third step. This step is charac-
terized by the rupture of one of two
Si-O-Si bridges, making the Q1Si
Na site.

Tamada et al.            Fig. 9
cessfully attack the last bridging oxygen to cleave the -2228. +
‐0.480
bridge or the sodium atom might dissolve out into the so-
‐0.485
lution. At this time, the computation needed to explore
Transition
ee)

this is too large. ‐0.490


0 490
y(Hartre

‐0.495
Promotion of basis set and correction of energies
Energy

‐0.500 88 kJ/mol
kJ/ l
Product
‐0.505
The energy barriers are as low as 22-29 kJ/mol in the sec-
‐0.510
ond and third steps. The highest barrier occurs in the first
step. The basis set was raised from the 3-21G* to the ‐0.515 Adsorption

6-311G(2d,p) level and applied to the transition and ad- ‐0.520

sorption states to better determine the energy barrier. The 1.5 2.0 2.5 3.0 3.5 4.0
Sit-Ow distance (Å)
barrier appreciably increased from the 63 kJ/mol calculat-
ed with the less robust basis to 88 kJ/mol as shown in Figure 10. The total energies of adsorption state, transition state
Figure 10. This increase is consistent with the increase and product re-optimized at the B3LYP/6-311G(2d,p) level start-
ing from their geometries in the first step. The energy barrier in-
Tamada et al.        Fig. 10
that was reported by Nangia and Garrison (2008), associ-
creased to 88 kJ/mol from the value of 63 kJ/mol calculated at
ated with the promotion of the basis set from 6-31+G(d,p) the B3LYP/3-21G* level.
to MG3S. Here we performed the vibrational frequency
calculation at room temperature. It is difficult to correctly
estimate the zero-point vibrational energy since positions the barrier is lowered to 85 kJ/mol. Not only zero point
of the matrix atoms were fixed in the optimization, but the energies (EZPE) of adsorption state and transition state,
rough estimation is reasonable because the zero-point vi- thermal energies (Etherm), thermal enthalpies (Htherm), entro-
brational energy itself is small compared with the total pies (S) and thermal free energies (Gtherm) are also estimat-
energy. It is also true that the matrix atoms are frozen ed from the vibrational frequency calculation as listed in
throughout the optimizations for the adsorption state and Table 2. Activation energy in enthalpy (ΔH) reduced to 82
the transition one, so that the effect on vibrational ener- kJ/mol and that in free energy (ΔG) increased to 91 kJ/
gies should be nearly equal. The energy difference for this mol. The experimental activation energy is best interpret-
part might, therefore, be cancelled out. The correction ed as the enthalpy of activation (∆H), i.e. the difference
value used for the activation energy is 3 kJ/mol, so that between the enthalpy of formation of the activated com-
94 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

Table 2. Activation energies Ea for the silica dissolution with change but the coordination number of the silicon atom
NaOH and without NaOH increases from four to five. The energy barrier of this re-
action is 63 kJ/mol in the B3LYP/3-21G* level. The dis-
tances between Si and coordinated oxygen atoms are
lengthened from about 1.65 Å to 1.75 Å associated with
the increase of coordination number, which induces the
weakening of Si-O bonded interactions. This weakening
suggests that the Si-O connections are susceptible to rup-
ture. It is also important to note that the product is at the
stable state having the minimum energy configuration as
shown in Figure 2.
In each of the second and third steps a water mole-
cule approached the five-fold coordinated silicon atom
and cleaved one of the weakened Si-Obr bonded interac-
tions in each step, with low energy barriers of 22-29 kJ/
mol (due to the increase in coordination number). In this
process, the Q3Si site changed to a Q2Si and then to a Q1Si
site.
Going back to the first goal of this study, the most
Note; The systems in the first step, re-optimized at the B3LYP/6-
311G(2d,p) level. important factor of the dissolution kinetics in the presence
Zero point energies (EZPE), thermal energies (Etherm), thermal en- of sodium hydroxide is that sodium plays the role of
thalpies (Htherm), entropies (S) and thermal free energies (Gtherm) are forming a five-fold coordinated silicon atom, finally keep-
derived from the vibrational frequency calculation at room tem- ing the system at the energy minimum. This makes the
perature (25 °C).
rupture of the Si-O bonded interactions easier and faster.
This result suggests that alkali ions like Li and K and al-
plex and the enthalpy of formation of the reactants, so that kali earth ions like Mg and Ca behave in the same way as
activation energy in enthalpy (ΔH) 82 kJ/mol, is hereafter sodium. Thus, the weakening of the Si-O bonded interac-
used for the comparison with the experimental one. tions resulting from the five-fold coordination should be a
The same calculations were also applied to the pro- clue as to the reason for the alkali or alkaline earth en-
cess without NaOH in the first step and the results are hancement of the silica dissolution.
listed in Table 2. The barrier appreciably increased from It is well known that the silicon atom is four-fold co-
the 112 kJ/mol calculated with the less robust basis to 150 ordinated by oxygen atoms in ambient condition. But co-
kJ/mol in the same way as above. The activation energies ordination numbers higher than four have been actually
in enthalpy ∆H and in Gibbs free energy ∆G are 139 and reported in glasses containing alkali oxide in the system
151 kJ/mol, respectively, which all demonstrates high bar- of Na2O-SiO2-P2O5 (Dupree et al., 1988, 1989) where the
riers when the sodium hydroxide is not introduced in the proportion of six-coordinated silicon depends on the al-
system. The entropy term T∆S at the room temperature is kali content and cooling rate. However, six-coordinated
9 and 12 kJ/mol for the reactions with and without NaOH, silicon atoms were not observed by NMR spectroscopy in
respectively. Their difference is 3 kJ/mol, which is mostly SiO2-P2O5 glasses without Na2O (Weeding et al., 1985).
negligible. After these reports a number of examples of six-coordi-
nated silicon atoms in SiO2-P2O5 glasses containing alkali
DISCUSSION oxide have been reported. Recently Ide et al.(2007) stud-
ied R2O-SiO2-P2O5 (R = Li, Na, K) glasses and observed
Penta-coordination of Si and reaction barriers six-coordinated silicon atoms in the XANES and XAFS
analyses and the fraction of six-coordinated silicon atoms
In the first step in which a water molecule and a sodium increased with increasing concentration of alkali oxide.
hydroxide were introduced to the system, nothing drastic Thus, these experimental results support our computa-
like a ruptured Si-O bonded interaction occurred but the tional result that alkali ions help to make silicon atoms be
process of moving a water molecule steadily toward a tar- coordinated by more than four oxygen atoms.
get silicon concluded with a five-fold coordinated silicon Wallace et al. (2010) investigated the ability of M(II)
atom. This is the consequence of the presence of a sodium aqua ions to promote the hydrolysis of Si-O bonded inter-
hydroxide. The connectedness of the Q3Si site doesn’t actions using the silica cluster model H8Si6O16 and ad-
Silica dissolution catalyzed by NaOH 95

sorption complexes of Mg2+(6H2O) and Ca2+(6H2O). The one is that the last Si-O-Si bridge is ruptured with the low
activation free energies are 178.9 (Mg) and 178.8 (Ca) kJ/ activation energy of 20-30 kJ/mol after the addition of
mol in the outer-sphere adsorption hydrolysis and 138.9 several more water molecules, since the target Si is still
(Mg) and 156.4 (Ca) kJ/mol in the inner-sphere adsorp- coordinated by five oxygen atoms. The second scenario is
tion hydrolysis, which is clearly different from the experi- that the sodium ion dissolves out into the solution by the
mental activation energies and our energy barrier of 82 addition of water molecules and the energy for the hydro-
(Na) kJ/mol. lysis of last Si-O-Si bridge is the same as the energy in
The second goal for the activation energy is fairly the deionized water without alkali ions, which conforms
well addressed in this study. The largest energy barrier is with the experimental result. This second scenario also
not the energy to rupture the Si-O bonded interactions but conforms with Pelmenschikov’s proposal. Pelmenschikov
the energy to form the five-fold coordinated silicon atom et al. (2001) proposed the new mechanism based on the
in the presence of sodium hydroxide. The barrier is 63 kJ/ self-healing effect of reactions which makes the preexpo-
mol in the 3-21G* level and 88 kJ/mol in the 6-311G(2d,p) nential factor extremely small and emphasized the impor-
level. The value 88 kJ/mol (82 kJ/mol in enthalpy) is fair- tance that the measured activation energy is associated
ly high compared with experimental values but is still in with the hydrolysis of the last Si-O-Si bond of the Si at-
the range of experimental ones of 46-96 kJ/mol. The en- oms.
ergy barrier of the Si-O rupture for Q3Si to Q2Si was 22 According to the second scenario our results are es-
kJ/mol and that for Q2Si to Q1Si was 22 or 29 kJ/mol in sentially consistent with Pelmenshikov et al.’s “self-heal-
this study, which does not indicate the special resistance ing” model of silica dissolution and provide a plausible
of the bulk structure as Pelmenschikov et al. (2000) re- explanation for why silica dissolution in the presence of
ported. This can be explained by the result that the Si-O alkali and alkaline earth ions is 20 to 30 times faster than
bonded interactions are already weakened by the five-fold in pure water even though the observed activation energy
coordination of the silicon atom in the presence of sodium is essentially the same for both cases. Dove and Crerar
hydroxide, so that the effects of bulk resistance are not (1990) showed that the dissolution rate of quartz in 0.05
explicitly observed in this study. m NaCl solutions is about 30 times faster than in pure wa-
Cypryk and Aleloig (2002) calculated the activation ter even but activation energies for both cases are essen-
energies for hydrolysis of siloxanes and siloxanols using tially the same [Ea(NaCl) = 71.5 kJ/mol and Ea(pure wa-
water clusters (H2O)n n = 1, 2, 3, 4 at B3LYP/6-311+G(2d,p) ter) = 71.3 kJ/mol]. Similarly, Icenhower and Dove (2000)
level and found that the energies are reduced from 140 kJ/ reported that the dissolution rate of amorphous silica in
mol to 74 kJ/mol as the number of water molecules in- 0.05 m NaCl solutions is about 20 times faster than it is in
creases, assisted by the cooperative proton transfer from pure water but the activation energies are nearly the same
the nucleophile to siloxane oxygen. This is the result in [Ea(NaCl) = 74.5 kJ/mol and Ea(pure water) = 76.4 kJ/
cleaving an Si-O-Si bridge of connectedness (Q1 → Q0) mol]. Thus, it seems that Na produces a catalytic effect
attacked by multiple water molecules. Although the situa- even though it does not change the experimental activa-
tion is quite different from our case, their result leaves the tion energy. A reasonable explanation for this effect is that
possibility of a reduction of the activation energy by the Na does not catalyze the Q1 → Q0 detachment reaction,
attack of multiple water molecules (n>3) on one Si-O-Si which is the observable step for silica dissolution, but
bridge, so that the present value 88 kJ/mol (82 kJ/mol in rather it catalyzes prior bond breaking steps in the overall
enthalpy) should be the upper limit of the activation ener- dissolution process such that it increases the concentration
gy. of Q1Si sites on the silica surface.
The release of dissolved Si from a silica surface in-
Dissolution associated with Pelmenschikov`s self- volves a chain reaction consisting of three bond breaking
healing effect steps:

After the third step the target silicon is still five-fold coor-
dinated and left as a Q1Si. The fourth and fifth steps were  
tried adding a water molecule in each case, but the trial (1)
failed because the last Si-O-Si bridge formed a narrow
bottle neck connecting the surface group and the inner where the overall rate of dissolution is
bulk structure, and the sodium prevented the water mole-
  r = k+1aH OCQ1 −k−1mH SiO (2)
cule from entering the bottle neck. In this situation two 2 4 4

scenarios seem plausible for the Q1Si rupture. The first If the solution is very dilute so mH SiO ≈ 0 and aH O =1, the
4 4 2
96 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

observed rate is simply a product of the forward rate con- CONCLUSIONS


stant and the concentration of the Q1Si sites
It is found that the presence of sodium atoms plays a role
  r = k+1CQ (3) for the formation of five-fold coordinated silicon atoms in
1
the dissolution kinetics of silica, which induces the elon-
at steady state gation of the Si-O distances and the weakening of the
bonded interactions, making the Si-O rupture easier and
 k+2CQ −k−2CQ =k+1CQ −k−1mH SiO (4) faster. Accordingly, the energy for the rupture of Si-O
2 1 1 4 4
bonded interactions becomes as low as 22-29 kJ/mol, but
the concentration of Q1Si sites depends on the concentra- the energy for forming the five-fold coordinated silicon
tion of Q2Si sites atom is substantially higher at 63 kJ/mol in 3-21G* level
and 88 kJ/mol (82 kJ/mol in enthalpy) in 6-311G(2d,p)
level, which is consistent with the range of experimental
(5)
   values. This is the example to report the lowest energy
barrier based on transition state theory for the Q3Si →
The concentration of Q2Si sites depends on the concentra- Q1Si hydrolysis in a neutral surface condition.
tion of Q3Si sites in a similar fashion. This chain of re- The calculations in this study, suggest a new under-
versible elementary reaction steps would produce the standing of the mechanism whereby an alkali cation en-
“self-healing” effect described by Pelmenschikov et al. hances the silica dissolution rate. This is based on increas-
(2001), which means that the concentration of Q2Si and ing the coordination number of the silicon atom from four
Q1Si sites is controlled by a sequence of near equilibrium to five. The process of the Si-O hydrolysis occurs as the
reaction steps. This leads to the conclusion that the mea- water molecule attacks the five-fold coordinated silicon
sured activation energy for silica dissolution simply re- through the intermediate of six-fold coordinated Si. The
flects the Q1Si to H4SiO4 reaction and that the overall dis- last rupture of Q1 to Q0 still remains for investigation.
solution rate is low because the concentration of Q1Si
sites is low. REFERENCES
The model presented here shows that inner-sphere
Bartelmehs, K.L., Gibbs, G.V., Boisen Jr., M.B. and Downs, R.T.
sorption of Na causes a significant reduction of the activa- (1993) Interactive computer software used in teaching and
tion energy barriers for the Q3Si to Q2Si and Q2Si to Q1Si research in mineralogy at Virginia Tech. Abstract with Pro-
reactions. This in turn causes the rates of these reactions grams of Geological Society of America Fall Meeting, the
to increase and increases the concentration of Q1Si sites, Geological Society of America, Boston, A-347.
which causes an increase in the overall reaction rate. Bennett, P.C. (1991) Quartz dissolution in organic-rich aqueous
systems. Geochimica et Cosmochimica Acta, 55, 1781-1797.
However, our model shows that the Q1Si to H4SiO4 reac- Brady, P. and Walther, J. (1990) Kinetics of quartz dissolution at
tion does not occur as long as the Na remains on the sur- low temperatures. Chemical Geology, 82, 253-264.
face but we postulate that Na will eventually become hy- Burnham C.W. (1997) Magmas and hydrothermal fluids. In Geo-
drated and depart the surface leaving behind a Q1Si site as chemistry of hydrothermal ore deposits 3rd edition (Barnes, H.
indicated above in the second scenario, which then reacts L. Ed.). pp. 972, John Wiley and Sons, New York, 63-123.
Cheng, H.P., Barnett, R.N. and Landman, U. (2002) Structure, col-
with a water molecule to produce H4SiO4. Because the lective hydrogen transfer, and formation of Si(OH)4 in SiO2-
more abundant Q1Si sites produced by this Na-catalyzed (H2O)n clusters. The Journal of Chemical Physics, 116, 9300-
process are indistinguishable from Q1Si sites produced by 9304.
the water hydrolysis elementary steps in the absence of Criscenti, L.J., Kubicki, J.D. and Brantley, S.L. (2006) Silicate
Na, the observed rate is increased even though the ob- glass and mineral dissolution: Calculated reaction paths and
activation energies for hydrolysis of a Q3Si by H3O+ using ab
served experimental activation energy for the Na-cata- initio method. The Journal of Physical Chemistry, 110, 198-
lyzed reaction is the same as for the reaction in pure wa- 206.
ter. Cypryk, M. and Apeloig, Y. (2002) Mechanism of the acid-cata-
This is the dissolution kinetics according to the lyzed Si-O bond cleavage in siloxanes and siloxanols: A the-
Pelmenschikov`s self-healing effect (Pelmenschikov et oretical study. Organometallics, 21, 2165-2175.
Dove, P.M. and Crerar, D.A. (1990) Kinetics of quartz dissolution
al., 2000, 2001) and all of these discussions emphasizes in electrolyte solutions using a hydrothermal mixed flow re-
the importance for determining the kinetics and the ener- actor. Geochimica et Cosmochimica Acta, 54, 955-969.
gy of hydrolysis for the last Si-O-Si bridge. Dove, P.M. and Rimstidt, J.D. (1994) Silica-water interaction. In
Silica: Physical behavior, geochemistry and materials appli-
cations (Heaney, P.J., Prewitt, C.T. and Gibbs, G.V. Eds.). pp.
Silica dissolution catalyzed by NaOH 97

606, Reviews in Mineralogy, 29, Mineralogical Society of 295.


America, Washington, D.C., 259-308. Lasaga, A.C. (1995) Fundamental approaches in describing min-
Dove, P.M. (1995) Kinetic and thermodynamic controls in silica eral dissolution and precipitation rates. In Chemical weather-
reactivity in weathering environments. In Chemical weather- ing rates of silicate minerals (White, A.F. and Brantley, S.L.
ing rates of silicate minerals (White, A.F. and Brantley, S.L. Eds.). pp. 583, Reviews in Mineralogy, 31, Mineralogical
Eds.). pp. 583, Review in Mineralogy, 31, Mineralogical So- Society of America, Washington, D.C., 23-86.
ciety of America, Washington D.C., 235-290. Ma, Y., Foster, A.S. and Nieminen, R.M. (2005) Reactions and
Dove, P.M. (1999) The dissolution kinetics of quartz in aqueous clustering of water with silica surface. The Journal of Chemi-
mixed cation solutions. Geochimica et Cosmochimica Acta, cal Physics, 122, 144709-1-9.
63, 3715-3727. Mazer, J.J. and Walther, J.V. (1994) Dissolution kinetics of silica
Du, M.H., Kolchin, A. and Cheng, H.P. (2003) Water-silica sur- glass as a function of pH between 40 and 85 °C. Journal of
face interaction: A combined quantum-classical molecular Non-Crystalline Solids, 170, 32-45.
dynamic study of energetics and reaction pathways. The Nangia, S. and Garrison, B.J. (2008) Reaction rates and dissolu-
Journal of Chemical Physics, 119, 6418-6422. tion mechanisms of quartz as a function of pH. The Journal
Dupree, R., Holland, D., Mortuza, M.G., Collins, J.A. and Locky- of Physical Chemistry, 112, 2027-2033.
er, M.W.G. (1988) An MAS NMR study of network-cation Nangia, S. and Garrison, B.J. (2009) Ab initio study of dissolution
coordination in phosphosilicate glasses. Journal of Non- and precipitation reactions from the edge, kink, and terrace
Crystalline Solids, 106, 403-407. sites of quartz as a function of pH. Molecular Physics, 107,
Dupree, R., Holland, D., Mortuza, M.G., Collins, J.A. and Locky- 831-843.
er, M.W.G. (1989) Magic angle spinning NMR of alkali Parks, G.A. (1984) Surface and interfacial free energies of quartz.
phospho-alumino-silicate glasses. Journal of Non-Crystalline Journal of Geophysical Research, 89, 3997-4008.
Solids, 112, 111-119. Pelmenschikov, A., Strandh, H., Pettersson, L.G.M. and Leszczyn-
Felipe, M.A., Xiao, Y. and Kubicki, J.D. (2001) Molecular orbital ski, J. (2000) Lattice resistance to hydrolysis of Si-O-Si
modeling and transition state theory in geochemistry. In Mo- bonds of silicate minerals: Ab initio calculations of a single
lecular modeling theory: Application in the geosciences (Cy- water attack onto the (001) and (111) β-cristobalite surfaces.
gan, R.T. and Kubicki, J.D. Eds.). pp. 531, Reviews in Min- The Journal of Physical Chemistry, B104, 5779-5783.
eralogy and Geochemistry, 42, Mineralogical Society of Pelmenschikov, A., Leszczynski, J. and Pettersson, L.G.M. (2001)
America, Washington, D.C., 485-531. Mechanism of dissolution of neutral silica surfaces: Includ-
Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, ing effect of self-healing. The Journal of Physical Chemistry,
M.A., Cheeseman, J.R., Montgomery, J.A. Jr., Vreven, T., A105, 9528-9532.
Kudin, K.N., Burant, J.C., Millam, J.M., Iyengar, S.S., To- Reed, M. H. (1997) Hydrothermal alteration and its relationship to
masi, J., Barone, V., Mennucci, B., Cossi, M., Scalmani, G., ore fluid composition. In Geochemistry of hydrothermal ore
Rega, N., Petersson, G.A., Nakatsuji, H., Hada, M., Ehara, deposits 3rd edition (Barnes, H.L. Ed.). pp. 972, John Wiley
M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakaji- and Sons, New York, 303-366.
ma, T., Honda, Y., Kitao, O., Nakai, H., Klene, M., Li, X., Rignanese, G.M., Charliar, J.C. and Gonze, X. (2004) First-prin-
Knox, J.E., Hratchian, H.P., Cross, J.B., Adamo, C., Jaramil- ciples molecular-dynamics investigation of the hydration
lo, J., Gomperts, R., Stratmann, R.E., Yazyev, O., Austin, mechanisms of the (0001) α-quartz surface. Physical Chem-
A.J., Cammi, R., Pomelli, C., Ochterski, J.W., Ayala, P.Y., istry Chemical Physics, 6, 1920-1925.
Morokuma, K., Voth, G.A., Salvador, P., Dannenberg, J.J., Rimstidt, J.D. and Barnes, H.L. (1980) The kinetics of silica-water
Zakrzewski, V.G., Dapprich, S., Daniels, A.D., Strain, M.C., reactions. Geochimica et Cosmochimica Acta, 44, 1683-
Farkas, O., Malick, D.K., Rabuck, A.D., Raghavachari, K., 1699.
Foresman, J.B., Ortiz, J.V., Cui, Q., Baboul, A.G., Clifford, Spear F.S. (1993) Metamorphic phase equilibria and pressure-
S., Cioslowski, J., Stefanov, B.B., Liu, G., Liashenko, A., temperature-time paths. pp. 799, Mineralogical Society of
Piskorz, P., Komaromi, I., Martin, R.L., Fox, D.J., Keith, T., America, Washington D.C..
Al-Laham, M.A., Peng, C.Y., Nanayakkara, A., Challa- Sverjensky, D.A. (2005) Prediction of surface charge on oxides in
combe, M., Gill, P.M.W., Johnson, B., Chen, W., Wong, salt solutions: Revisions for 1:1 (M+L−) electrolytes. Geochi-
M.W., Gonzalez, C. and Pople, J.A. (2003) Gaussian 03, Re- mica et Cosmochimica Acta, 69, 225-257.
vision B.04. Gaussian, Inc., Pittsburgh PA. Tester, J.W., Worley, W.G., Robinson, B.A., Grigsby, C.O. and
House, W.A. and Hickinbotham, L.A. (1992) Dissolution kinetics Feerer, J.L. (1994) Correlating quartz dissolution kinetics in
of silica between 5-35 °C. Journal of the Chemical Society, pure water from 25 to 625 °C. Geochimica et Cosmochimica
Faraday Transactions, 88, 2021-2026. Acta, 58, 2407-2420.
Ide, J., Ozutsumi, K., Kageyama, H., Handa, K. and Umesaki, N. Tullis, J. and Yund, R.A. (1980) Hydrolytic weakening of experi-
(2007) Journal of Non-Crystalline Solids, 353, 1966-1969. mentally deformed Westerly granite and Hale albite rock.
Iler, R. (1979) The Chemistry of Silica. pp. 866, John Wiley and Journal of Structural Geology, 2, 439-451.
Sons, New York. Tullis, J. and Yund, R.A. (1989) Hydrolytic weakening of quartz
Icenhower, J.P. and Dove, P.M. (2000) The dissolution kinetics of aggregates: The effects of water and pressure on recovery.
amorphous silica into sodium chloride solutions: Effect of Geophysical Research Letters, 16, 1343-1346.
temperature and ionic strength. Geochimica et Cosmochimi- Wallace, A.F., Gibbs, G.V. and Dove, P.M. (2010) Influence of
ca Acta, 64, 4193-4203. ion-associated water on the hydrolysis of Si-O bonded inter-
Lasaga, A.C. and Gibbs, G.V. (1990) Ab-initio quantum mechani- actions. The Journal of Physical Chemistry, A114, 2534-
cal calculations of water-rock interactions: Adsorption and 2542.
hydrolysis reactions. American Journal of Science, 290, 263- Walsh, R.T., Wilson, M. and Sutton, P.A. (2000) Hydrolysis of the
98 O. Tamada, G.V. Gibbs, M.B. Boisen Jr. and J.D. Rimstidt

amorphous silica surface: II. Calculation of activation barri- Xiao, Y. and Lasaga, A.C. (1994) Ab initio quantum mechanical
ers and mechamisms. The Journal of Chemical Physics, 113, studies of the kinetics and mechanisms of silicate dissolution:
9191-9201. H+(H3O+) catalysis. Geochimica et Cosmochimica Acta, 58,
Weeding, T.L., De Jong, B.H.W.S., Veeman, W.S. and Aitken, 5379-5400.
B.G. (1985) Silicon coordination changes from 4-fold to Xiao, Y. and Lasaga, A.C. (1996) Ab initio quantum mechanical
6-fold on devitrification of silicon phosphate glass. Nature, studies of the kinetics and mechanisms of quartz dissolution:
318, 352-353. OH− catalysis. Geochimica et Cosmochimica Acta, 60, 2283-
White, A.F. and Brantley, S.L. (1995) Chemical weathering rates 2295.
of silicate minerals. pp. 583, Review in Mineralogy, 31, Min-
eralogical Society of America, Washington D.C.. Manuscript received September 9, 2011
Worley, W.G., Tester, J.W. and Grigsby, C.O. (1996) Quartz disso- Manuscript accepted January 20, 2012
lution kinetics from 100-200 °C as a function of pH and ionic Published online April 7, 2012
strength. AIChE Journal, 42, 3442-3457 Manuscript handled by Hiroki Okudera

You might also like