Você está na página 1de 39

Silica I integral text 2/22/07

Cahiers of Geochemistry
Silica I: Silicon
Analytical, Physical and Terrestrial
Geochemistry
C.H. van der Weijden
Department of Geosciences Geochemistry
Utrecht University
P.O. 80.021
3508 TA Utrecht
The Netherlands
chvdw@geo.uu.nl
cornelisvdw@hotmail.com

Content:
I: Introduction
II: Analytical chemistry
III: Structural chemistry
IV: Physical chemistry
V: Kinetics of silica dissolution and(or) precipitation
VI: Thermodynamic properties
VII: Hydrothermal systems
VIII:Natural concentration
IX: Controls on dissolved silica concentrations
a) Role of vegetation
b) Role of chemical weathering
X: Lithification and silification
XI: References

I. Introduction
The crustal abundance of silicon is 28%, only surpassed by oxygen which makes up
46% of the crustal composition. Its naturally occurring isotopes are: 28Si (92.23%),

1
Silica I integral text 2/22/07

29
Si (4.67%) and 30Si (3.1%). Almost all silicon is present in silica (SiO2) polymorphs
and in (alumino)silicate minerals, very little as dissolved species in natural waters. As
a non-volatile element, silicon in the atmosphere is in wind-blown particulates, not as
a gaseous component. In the geochemical cycle, silicon is transported to the sea as
part of the solid and dissolved loads of rivers and as atmospheric dust. Although its
contribution in this cycle is small, dissolved silicon plays a crucial role in
biogeochemical cycles. Silicic acid is transformed into amorphous silica in
unicellular organisms, in silicon sponges, in plants such as grasses, reed, horsetail and
nettle plants (Kaim and Schwederski, 1994). The concentrations of silicic acid in
surface water and groundwater are usually low; in hydrothermal waters its
concentration is generally higher.

II. Analytical chemistry


Modern analytical techniques, such as X-ray Fluorescence Spectrometry (XRF) and
Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES), allow the
precise and accurate determination of total silicon in solid materials. For XRF
analysis, the solid sample is usually dissolved in a borate melt, poured into a mould,
allowed to cool, after which silicon is measured (K12 = 0.71250.7128 nm) in the so-
obtained glass disc using silicon in borate glass discs prepared from international
standards for calibration. For ICPOES analysis, the sample is first fused with sodium
hydroxide or carbonate, after cooling of the melt the solid material is dissolved in
water and silicon in the solution is measured ( = 288.16 nm) against solutions of
international standards prepared in the same manner.
More problematic is the determination of amorphous (am-SiO2) and opaline
(SiO2nH2O, also amorphous) silica.1 One method uses X-ray diffractrometry to
measure the opaline silica (Leinen, 1985). Briefly, first calcium carbonate (CaCO3)
when presentis removed by treatment with a sodium acetate buffer solution, then
the sample is thoroughly ground and alumina (-Al2O3) is added for use as an internal
standard, next this mixture is ignited at 1000C which converts opal into cristobalite
(-SiO2), and finally the peak areas at 0.348 nm (alumina) and 0.409 nm
(cristobalite), both corrected for background, are measured and compared with
standards that have been treated in the same way to determine the original silica
content. There are more caveats to this method than mentioned here (Leinen, 1985),
according to Mortlock and Froelich (1989) the method works satisfactorily, but
Conley (1998) shows that a systematic overrating of the biogenic silica (SiB) content
relative to wet chemical methods can occur.
More often, extraction of BSi in alkaline (NaOH or Na2CO3) solutions is used. A
complication is that not only BSi dissolves but that also clay minerals are attacked in
alkaline solutions. Eggiman et al. (1980) tested extraction of BSi in sediments in a
range of Na2CO3 concentrations and recommended the use of 2 M Na2CO3 at 90
100C for 4 hours in combination with correction for the contribution by clays using
the co-extracted Al (measured by AAS or ICP). Their correction factors depend on the
ratio between BSi and clay contents. Another methodological line uses differences in
rates of solution between opaline silica and other Si-bearing solids. When the
dissolved Si concentrations are plotted as a function of time (t), the initial
concentrations will rapidly increase parabolically as a result of BSi dissolution and

1
For brevity, the descriptions opal (opaline silica, biogenic opal) or BSi (biogenic silica) are
alternatively used instead of (biogenic) amorphous silica.

2
Silica I integral text 2/22/07

after some time (depending on T and pH) will continue to increase linearly due to the
dissolution of clays in the sediment matrix. The basic idea is that dissolution of BSi is
complete after a relatively short time and that clay minerals dissolve much slower.
The original BSi content is then estimated from the intersection of the extrapolated
linear part of the dissolution curve to t = 0. The error made by this method is
understandingly higher in sediments with a low than a high BSi content. The leaching
of Si from clay minerals in the sediment matrix decreases in the order
montmorillonite(smectite) > illite > kaolinite > chlorite (Mller and Schneider, 1993).
But interference from co-leached clay minerals is in general not a serious problem.
The composition of alkaline leaching solutions tested in well-known marine
laboratories varies from 0.1 to 2 mol/L Na2CO3 or NaOH (Conley, 1998). Actually,
Schlter and Rickert (1998) recommend a pH of 12.5 of the leaching solution. The
mostly used temperature is 85C and the leaching time varies between a few to eight
hours. Mller and Schneider (1993) and Koning et al. (1997) use automated methods
in which sampling and analysis of dissolved silicic acid is done within consecutive
short time intervals during a continuous extraction procedure. They give recipes for
the calculation of the BSi content from linear extrapolation (Mller and Schneider,
1993) and curve fitting (Koning et al., 1997) of the dissolved silica concentrations
versus leaching time. The liberated silicic acid (H4SiO40) is measured
spectrofotometrically (see hereafter). Thorough tests of these methods proved that
errors, introduced by simultaneous dissolution of clay minerals, is low. Mller and
Schneider (1993) estimate that the accuracy of their method is 104% for samples
with 210% BSi respectively and decreasing to 2% for opal-rich samples. To convert
the measured opaline silicon into opaline silica content, multiplication with a factor
2.4 can be used. According to Mortlock and Froelich (1989), opal has about 10% H2O
(SiO20.4H2O), so the opal content is 1.12SiO2 or 2.4Si.
Dissolved silicon in fresh water, seawater and in solutions obtained by dissolution of
silica and silicates is determined spectrofotometrically. Briefly, this method is based
on the reaction between silicic acid and ammonium (hepta)molybdate (tetrahydrate) in
an acidified solution (pH < 2.5) to form a yellow polymer. Since the molar
absorptivity of this polymer is low, one usually measures the reduced blue complex
obtained by addition of oxalic acid (to prevent reduction of excess molybdate and
interference with phosphate) and ascorbic acid. After complete evolvement of the blue
color, the absorption of the solution is measured. This method is often used in an
automated mode of sample introduction and absorption measurements in Auto
Analyzer apparatus which can also be applied on board of research vessels (Grasshoff
et al.,1983). Using this method, only the so-called molybdate-reactive silica is
measured, encompassing monomeric, dimeric and trimeric (Icopini et al., 2005)
dissolved silicic acid species. Total dissolved2 Si, including colloidal species, can be
measured by ICP-AES.

2
For samples of natural waters, dissolved Si is operationally defined by all species that pass the filter
pores (frequently used are filters with pore sizes of 0.45 or 0.2 m). It is of course possible to use
ultrafiltration techniques to achieve a better separation between true and colloidal dissolved Si. One
should be aware that storage of water samples by freezing prior to analysis may lead to immobilization
of dissolved Si (Tallberg et al. 1997).

3
Silica I integral text 2/22/07

III. Structural chemistry


Perry (1999) summarized some structural aspects of silicon oxides and her summary
will be used in this section. The SiO4 tetrahedron with Si in the center of this unit and
Os in contact with one another is the building bloc of most silicon containing
minerals, with the exception of SiO6 octahedral units in the high temperature minerals
stishovite and coesite. For all structures the SiO bond is about 0.162 nm, smaller
than the sum (0.191 nm) of the covalent radii of Si and O atoms. This accounts for the
high stability of the siloxane (SiOSi) bond. The SiO4 tetrahedra tend to form three-
dimensional framework structures in silica crystal chemistry.
Amorphous silica is also built up from SiO4 tetrahedra but with variable SiOSi
bond angles and SiO bond distances In contrast to their crystalline analogues,
amorphous silicas exhibit no long-range order, are known for their tendency to
produce disordered aggregates and exhibit a wide range of porosities depending on
particle sizes and aggregate structure. The concentration of surface silanol (SiOH)
groups in amorphous silica decreases with increasing pH, suggesting that silanol sites
polymerize to form siloxane bonds (cf. eq. SiI-14) with increasing pH (Carroll et al.,
2002).
Because of their high solubility, amorphous silicates do not prevail in biological
structures. Only when an unusual high energetic barrier exists, such as in SiO2.nH2O,
can this be overcome. Amorphous silica encompasses an almost infinite variety of
structural forms, from hydrated to gel-like materials. In all these cases, amorphous
silica exists as a hydrated, covalent inorganic polymer with the general formula
(SiOn/2(OH)4-n)m, in which n = 04 and m is a large number. A similar variation
applies for the extent of hydration.

IV. Physical chemistry


The following dissolution/precipitation and dissociation equilibria have to be
considered at/near the Earths surface:

SiO2(s) + 2 H2O H4SiO40 pKsp = 4 (-quartz) (eq.SiI-1)


monomeric silicic acid pKsp = 2.7 (opaline silica) (eq.SiI-2)
H4SiO40 H3SiO4 + H+ pK1 = 9.82 (eq.SiI-3)

In the normal pH-range of natural waters (49), the major dissolved species is
H4SiO40 and at pH = 9.82 the activities of H4SiO40 and H3SiO4 will be equal.
Solutions in equilibrium with metastable amorphous silica are about twenty times
oversaturated with respect to quartz. Since recrystallization of amorphous silica into
more stable SiO2 polymorphs is a very slow process, supersaturation is not
uncommon in natural waters. At unusually high pHs the second dissociation step will
become important:

H3SiO4 H2SiO42+ H+ pK2 = 13.1 (eq.SiI-4)

The solubility of amorphous silica and quartz as a function of pH and the stability
fields of the monomeric species are shown in figure SiI-1.

4
Silica I integral text 2/22/07

Dimerization can be expected in solutions high in dissolved silicon

2 H4SiO40 H6Si2O70 + H2O pKdim = 2.5 (eq.SiI-5)

In high ionic strength (I) solutions comparable to the I of seawater, the apparent
dissociation constants are given by

H 0
K1* = K1 ( 4 SiO4
) (eq.SiI-6a)
H H SiO
+
3 4

H SiO
K 2* = K 2 ( 3 4
) (eq.SiI-6b)
H H SiO
+ 2
2 4

At I = 0.723 molal (average seawater), H + = 0.82, H 0 = 1.167, H = 0.624,


4 SiO 4 3 SiO 4

and H
*
2 = 0.123, which means that in seawater pK1* = 8.83 and pK2 = 12.31
2 SiO4

(Hershey and Millero, 1986). The authors give values of the s for 0.05 I (molal) <
2, calculated by applying Pitzers equations.

Fig. SiI-1: Solubility of quartz and amorphous silica as a function of pH; the solubility
at each pH value represents the sum of the dissolved silicon species as
indicated (redrawn from Hershey and Millero, 1986).

Tossell (2005) used a quantum mechanical approach to estimate the equilibrium


constant for dimerization and arrived at a value of ~101.5. At elevated T and P, with a
strongly decreased dieletric constant, the dimerization will become more favorable.

5
Silica I integral text 2/22/07

At the level of most freshwater and seawater Si concentrations, monomeric silicic


acid is, for all practical purposes, the only species to be considered (Isshiki et al.,
1991; Dietzel, 2000).

The dependency of silica solubility on ionic strength (I) is given by the Setchenow
equation (Langmuir, 1997):

log H 0 = 0.0803I (eq.SiI-7, but cf. with eq. SiI-10b)


4 SiO 4

For seawater (I 0.7) the decrease in solubility amounts to 14% compared to the ideal
system. The apparent first dissociation constant of silicic acid in seawater is given by
(Hershey and Millero, 1986)3:

K1* = 109.47 = 1.89 [H+](H3SiO4)/(H4SiO40) (eq.SiI-8)

The solubility of silica in brines depends not only on I, but also on the composition of
the solution. Since [H2O] < 1 at high I, the solubility of silica can be represented by

(H 4SiO 4 ) = [H 2O] K sp / H SiO 0 (eq.SiI-9)


4 4

When silicic acid concentrations are plotted against water activities, different
electrolytes show different effects (Figure SiI-2)

3
[X] is used for activity and (X) for concentration of species X

6
Silica I integral text 2/22/07

Fig. SiI-2: Effect of decreasing water activity on the solubility of amorphous silica for
solutions of 1:1 and 1:2 salts (redrawn from Iler, 1979)

For salt solutions, the solubility is also related to the hydration numbers of the cations.
According to Weres et al. (1981) the solubility of amorphous silica can be expressed
as a function of the salt concentration with I 6 by

c o (m) = c o (0) aw2 e[0.018( n )m ] (eq. SiI-10b)

where co(m) is the solubility of the silica in a solution of the given salt of molality4 m,
co(0) is the solubility at m = 0; co is estimated by the empirical formula of Fournier
and Rowe (1977): logco = 731/T + 1.52 (T in K); aw is the water activity of the
solution, is the number of ions per unit formula of the salt, and n is the apparent
hydration number for a given salt.

4
Note that molality is defined as mol/kg H2O and molarity as mol/liter of solution

7
Silica I integral text 2/22/07

Fig.3: Effect of the hydration number of cations on the solubility of amorphous silica
for a fixed molarity of salt solutions (redrawn from Iler, 1979)).

Precipitation of silica from oversaturated solutions proceeds by condensation


reactions, either by polymerization or by adsorption onto existing silica (or other
suitable surfaces). The following steps can be distinguished (Weres et al., 1981): 1)
formation of silica polymers of less than critical size, 2) nucleation of an amorphous
silica phase in the form of colloidal particles, and 3) growth of supercritical
amorphous silica by further chemical deposition of silicic acid on their surfaces.
Dissolved silica may also be deposited onto solid surfaces other than silica colloids by
analogy of step 3, which is called molecular deposition. Step 1 is usually fast in
comparison with step 2, and therefore step 2 is rate determining. The polymerization
reactions can be expected when the silicon concentration exceeds the solubility of
amorphous silica (~ 2 mM or 120 mg L1).

8
Silica I integral text 2/22/07

Oligomerization5 precedes polymerization. Icopini et al (2005) derived the following


rate of formation of molybdate-reactive silica, (SiO2)n3:

R = k4[H4SiO40]4 (eq. SiI-11c)

with

logk4 (mmolal3s1) = mpH + logk0 (eq. SiI-11b)

Where k0 is the rate constant at pH 0 and m is the slope of the logR versus pH curve;
m is positive for pH < 7 and negative for pH > 7.

Icopini et al. (2005) concluded that the observed fourth-order rate dependence with
respect to [SiO2]n3 is consistent with a critical nucleus of four monomers as a cyclic
tetramer. With regard to the presence of salts, such as present in brines, Weres et al.
(1981) found that the more a salt depresses the solubility of amorphous silica (cf. eq.
SiI-10b), the more it will accelerate the polymerization rate. Small concentrations of
fluoride play a distinct role in the polymerization rate. According to Weres et al.
(1981), this rate depends on the total F concentration as well as on F speciation (HF:
pKa = 3.2). Aluminium attached to or incorporated in the silica surface reduces the
silica solubility (Iler, 1973) and has an inhibitory effect on the molecular deposition
too (Weres et al., 1981).
Polymerization can be described as the mutual condensation of silanol groups to give
molecularly coherent units of increasing size, whether these are spherical particles of
increasing diameter or aggregates of an increasing number of constituent particles
(Iler, 1979). The basic reaction is

SiOH + HOSi SiOSi + H 2O (eq.SiI-12)


silanol groups siloxane group

For pH > 2, the rate of polymerization turns out to be proportional to pOH (= pKw
pH). This can be represented by the following reaction mechanism

SiOH SiO + H+ (eq.SiI-13)


+
H + OH
H 2O (eq.SiI-14)
-------------------------------------------+
SiOH + OH SiO + H2O (eq.SiI-15)
SiOH + SiO SiOSi + OH (eq.SiI-16)
-------------------------------------------+
2 SiOH SiOSi + H2O (eq.SiI-17)

The condensation process tends to produce a maximum of siloxane and a minimum of


silanol groups, proceeding from formation of ring structures via linking together the
cyclic polymers to form 3D molecules. The internal silanol groups condense to the
most compact state with silanol groups remaining on the outside. The solubility of the
spherical particles depends on the particle size and on the completeness of
dehydration of the internal solid phase (below 80C, internal silanol groups may still

5
According to Dietzel (2000), oligomeric or low-molecular-weight silica is used up to decamer silica,
and polymeric or high-MW silica for polysilicic (> 10 Si) acids.

9
Silica I integral text 2/22/07

be present). Larger particles grow at the expense of small ones (Ostwald ripening).
Typically, particles with diameters of 510 nm (for pH > 7) and 24 nm (for pH < 7)
increase only very slowly in size, but further growth will be enhanced by a raise of
temperature. In the range of 7 < pH < 10.5, the particles carry a large negative charge
and will repel each other, so particle growth continues by incorporation of monomeric
silicic acid. In the presence of salts, however the surface charge becomes reduced and
aggregation and gelling occurs. At pH < 7, the silica particles have a very low charge
and can collide, aggregate and form gel networks (Iler, 1979). Van Cappellen et al.
(2002) argue that at pH > 4 the surface of biogenic silica has only SiOH0 and
SiO groups and that a negative surface charge builds up rapidly at pHs > 6.

The suite of condensation reactions, proceeding via dimers, oligomers, polymers and
finally to solid silica, can be represented as follows:

OH
(HO)3SiOH + m(HOSiOH) + HOSi(OH)3 + HOSi(OH)3
HO

OH
(HO)3SiO(SiO)mSi(OH)3 + HOSi(OH)3 + mH2O (SiOm(OH)2m+8)s (eq.SiI-18)
HO amorphous silica

where m > 1.

Assuming that amorphous silica exists as discrete particles with anhydrous SiO2 in the
interior, the number of surface silanol groups depends on the value of m. The smaller
a particle, the higher the number of silanol groups; this is illustrated in Table SiI-1

n diameter (in nm) OH:Si O:Si


anhydrous basis
8 0.89 0.99 1.51
100 1.52 0.85 1.57
1438 5 0.37 1.82
11500 10 0.2 1.9
0 2

Table SiI-1: ratios of hydroxide and oxide groups relative to silicon as a function of
the number (n) of silicon atoms in spherical particles of amorphous silica
(Marshall and Warakomski, 1980).

The intrinsic acidity of silanol groups on the surface of silica or polysilicic acid is
higher (pKintr 6.8) than of monomeric silicic acid (pK = 9.82). The dissociation
reaction of silanol and the concomitant dissociation constant are

SiOH SiO + H+ pK = pH + log(1-)/ (eq.SiI-19)


(1-)

where is the fraction of silanol groups that are ionized at a given pH.

10
Silica I integral text 2/22/07

Based on experimental results, a relation between pH and is represented by (Iler,


1979):

pH = pKintr log(1-)/ (0.039+)/1.9(+2) (eq.SiI-20; cf. with eq. SiI-19)

The iep (isoelectric point: pH at which there is no mobility of silica particles in an


electric field) and the pzc (zero point of charge: pH at which the net surface charge is
zero) for silica polymorphs is 2 0.5. This means that in the pH-range of natural
waters, silica surfaces have a negative charge and offer reactive sites for cation
sorption. A relation between the surface charge of quartz, (in the presence of only H+
and OH as charge determining ions) and pH is (Walther, 1996)

log(SiO) = 0.3 pH 12.67 (5 < pH < 12) (eq.SiI-21)

where (SiO) is the negative surface charge in moles cm2.

Ionic Al-species may adsorb onto silica surfaces or Al may even replace Si in a silica
surface. As mentioned before, these processes lead to a decrease of the rate of solution
as well as of the solubility of silica (Iler, 1973) even when less than a monolayer of Al
ions is adsorbed on the surface or when replacement of Si by Al is incomplete. Such
processes are held responsible for the gradual lowering of opal solubility in the
oceanic water column and, more pronounced, in marine sediments with a high
reactive-Al content. In natural waters the dominant dissolved Al-species are Al3+ (pH
< 5), Al(OH)2+ (pH 6) and Al(OH)4 (pH > 6.3) (Langmuir, 1997). In the marine
domain, pH > 6.3 so the negative Al-hydroxy species should be the incorporated in
the silica structure, which seems to be at odds with the negative surface charge of
biogenic silica in seawater and marine pore waters.
Spadini etal. (2004) studied complexation of Si and Al and derived for the reaction

Al3+ + H4SiO40 AlOSi(OH)32+ + H+ p1,1,1 = +2.30 (I 0) (eq. SiI-22)

or

Al3+ + SiO(OH)3 AlOSi(OH)32+ pK = 6.72 (I = 0.6 M) (eq. SiI-23)

The latter value means that in seawater the stability of the Al(III) complexes
decreases in the order OH (pK = 8.2) > SiO(OH)3 (pK = 6.72) > F (pK = 6.16)

Usually the dissolved Al concentrations in weathering environments are much lower


than the Si concentrations. Neoformation of kaolinite (0.5Al2O32SiO22H2O) will
control the Al concentration in the presence of excess silicic acid. The pertinent
reactions under acidic and basic conditions are

H4SiO40 + Al3+ + H2O kaolinite + 3 H+ logK = 3.72 (eq.SiI-24)


H4SiO40 + Al(OH)4 kaolinite + OH + 2.5 H2O logK = + 4.98 (eq.SiI-25)

Kaolinite is the most stable mineral phase, but according to the Ostwald step rule,
allophane and halloysite may be precursors of kaolinite in weathering sequences.

11
Silica I integral text 2/22/07

These phases have higher solubilities than kaolinite. The discussion on this subject
will be continued in the chapter on weathering.

V. Kinetics of silica dissolution and(or) precipitation


The rate of silica solution depends on pH. At pHpzc/iep= 2 0.5, protonation (pH <
pHpzc/iep) as well as deprotonation (pH > pHpzc/iep) enhances the dissolution rate.
The dissolution rates (far from equilibrium) of (amorphous) silica glass depend on
pH. Wirth and Gieskes (1979) observed a linear dependence between the logs of
dissolution rate and the absolute value of the surface charge density (), the latter in
turn depending on the pH. The slope of this plot is ~2, suggesting that there are two
surface excess hydroxides (or H+ deficiencies) in the transition state of the rate-
determining step of silica dissolution at lower surface charge. Plotted against pH or
logI (ionic strength), a linear correlation with log is found. The effect of I on the
pH-dependence of the surface charge was not very strong, but the slope of log
plotted against logI changes with pH. The effect of I will be discussed in more detail
in the last parts of this section. Mazer et al. (1994) studied the dissolution rates of
vitreous silica as a function of pH. They report that the rate first increases
parabolically ( t) and then turns into a linear ( t) rate. They argue that the
parabolic kinetics are due to diffusion of H2O into the glass structure, allowing a rapid
release of weakly bonded tetrahedra. Formulated otherwise, the initial rate of
diffusion of hydrated Si species out of the glass exceeds the rate of the surface
detachment of tetrahedral rings Si units. With time, the diffusion path length of
hydrated Si out and of H2O into the structure increases and the surface detachment of
the tetrahedral rings becomes rate controlling. The parabolic and linear rates with
respect to pH are similar: for pH < 4, the slope of the parabolic rate constant is ~
0.1, for pH < 7 the slope of the linear rate constant is ~ 0.15; for pH > 8, this slope is
~ +0.4 for the parabolic and ~ +0.5 for the linear part. In studies on the dissolution of
quartz, no initial parabolic dissolution is reported, which can be explained by the
difference in SiO bonding between vitreous silica and quartz (fig. SiI-4)

Recently, Carroll et al. (2002) investigated the dissolution rate of amorphous silica.
The overall rate could be fitted using the following equation (Fig. SiI-5)

( E a 1 / 2.303 RT )
rate(mol m2 s1 ) = 10 k1 10 + 10 k2 10( E a 2 / 2.303RT ) (eq. SiI-26)

where R is the gas constant, T the temperature in K, Eas are activation energies of
dissolution by hydrolysis and deprotonation:
Ea1=26.9 and Ea2 = 8.15pH (in kJ mol1), ks are concomitant rate constants:
k1 = 5.7 and k2 = 13.7 + 1.9pH (mol m2 s1).

12
Silica I integral text 2/22/07

Fig. SiI-4: Rates of dissolution of vitreous silica (Wirth and Gieskes, 1979) and quartz
(Wollast and Chou, 1985) as a function of pH (redrawn from Stumm,
1992).
.

13
Silica I integral text 2/22/07

Fig. SiI-5: Dissolution rate of SiO2 glass at 22C as a function of pH. The experiments
were conducted in electrolyte solutions of 0.011 molal NaCl and in 0.01
0.1 molal CsCl. The solid line represents the rates calculated by equation
SiI-26, the dashed line was visually fitted to the data points obtained in
experiments with 0.01 molal NaCl (redrawn from Carroll et al., 2002)

In the presence of alkali ions, the dissolution rates are a little lower, and can be fitted
(Fig. SiI-6) by the equation

lograte(molm2s1) = k1 + k2[SiO] + [SiOM+], (eq. SiI-27)

where k1 = 5.25 29.7/(2.303RT) and k2 = 107[1.06 5.3/(2.303RT)] (mol m2 s1).


The third term on the right-hand side is written as an outer sphere complex.

14
Silica I integral text 2/22/07

Fig. SiI-6: Dissolution rate of SiO2 glass at 22C as a function of the total ionized
surface sites. The experiments were conducted in electrolyte solutions of
0.011 molal NaCl and in 0.010.1 molal CsCl. The solid line represents
the rates calculated by equation SiI-27 (redrawn from Carroll et al., 2002)

It is important to compare laboratory with field results, because this is a test for the
applicability of experimental models outside the laboratory where conditions are
controlled and varied arbitrarily.
In weathering environments and hydrothermal systems, -quartz is the stable silica
phase. Under ambient conditions quartz has a low solubility, as is evident from the
fact that sand beaches often fringe the seaboard. However, quartz does have a finite
solubility (~ 6 ppm at 25C) andcontrary to the solubility of calcium carbonateits
solubility increases with temperature. Upon cooling, silica may precipitate from
hydrothermal solutions, resulting in quartz veins and scaling of tubes in hydrothermal
power plants.

In the pH-range of natural waters, the dissolution rate but not the solubility is pH
dependent, because solubility starts to increase considerably only from pH > pK12.
In the acid pH range, the mechanism controlling the rate of solution is the protonation
(H+ or H3O+) of oxygen in the SiOSi bridging bonds, significantly weakening
these bonds and facilitating the detachment of solvable molecular Si-units. In
comparison to the hydrolysis of SiOSi by H2O, the catalysis by H3O+ lowers the

15
Silica I integral text 2/22/07

activation energy of hydrolysis by some tens of percent (Xiao and Lasaga, 1994).
Under basic conditions, the catalytic mechanism of hydrolysis starts as an OH attack
on the hydroxyl (silanol) surface leading to deprotonation and resulting in a H-bonded
H2O adsorption on a negatively charged SiO surface; the next step is supposed to
involve the formation of a negatively charged fivefold coordinated Si species, a step
with the highest energy barrier, but this intermediate activated complex weakens the
SiO bond; the final step is the rupture of this bond to form SiOHOSi
(Xiao and Lasaga, 1996).

Apart from the pH dependence, the ionic strength (I) of the solution has an effect on
the dissolution rate of silica. An increase of I promotes deprotonation of surface sites
and subsquently the rate of solution. For 7 < pH < 10.5, 0.001 < I <1 and at 25C,
the following relation was derived for this I-dependent increase of negative surface
sites:

( log [SiO]/ log I)pH 0.2 (eq.SiI-28)

By applying this relation, observed dissolution rates at different Is could be


converted to rates at the same ionic strength (Brady and Walther, 1990).

The dissolution kinetics of quartz are also influenced by the solution chemistry. The
dissolution rate was found to be slowest in pure water and increasing in the presence
of single chloride salts (0.0001 to 0.2 m) in the order Mg2+ < Ca2+ Li+ Na+ K+ <
Ba2+ in experiments at near-neutral pH and 200C (Dove and Nix, 1997). This trend
compares well with results obtained at 25C. The catalytic effect of salts on quartz
dissolution decreases with increasing silicic acid concentration. There appears to be a
relation (Fig. SiI-7) between the frequency of solvent exchange, kex (s-1); this is the
rate at which primary solvation waters molecules, associated with the individual
cation, exchange with the surrounding aqueous environment (can be measured by
NMR on 17O exchange rates).

16
Silica I integral text 2/22/07

Figure SiI-7: Correlation of the rate of quartz solution (rdiss) and the frequency of
solvent exchange (kex) of salt solutions (redrawn after Dove and Nix,
1997))

The assumed mechanism of catalytic dissolution is via adsorption onto surface silanol
groups

SiOH + Men+ SiOMe(n1)+ + H+ (eq.SiI-29)

which weakens the SiOSi bonds and promotes hydrolysis. In mixtures of cation
solutions, the dissolution rates are a nonlinear combination of the bulk concentrations
of cations in such manner that the rates are limited by the cation with the smallest
rate-enhancing effect, e.g. Mg2+ (Dove, 1999). The explanation is that cation
adsorption is controlled by the ion potential (charge/radius of ion). A cation with a
high ion potential displaces one with a lower ion potential (e.g. Ba2+), so the surface

17
Silica I integral text 2/22/07

complexes are dominated by the cations having the highest ion potential. Hence, net
dissolution rates in solute mixtures are determined by the population of cations at the
surface.

The dissolution rate for the reaction quartz + water silicic acid, can be formulated
as (Rimstidt and Barnes, 1980)

r = ([H4SiO4]/t)P,T,M, = k+ k[H4SiO40] (eq.SiI-30)

where r = net rate of production of H4SiO40.


k+ = rate constant for dissolution.
k = rate constant for precipitation.

Considering that the ratio of the forward (dissolution) and backward (precipitation)
reaction rate constants is represented by the thermodynamic equilibrium constant, K,
one can derive (Rimstidt and Barnes, 1980)

r = k+(1 Q/K) (eq.SiI-31)

where Q and K = [H4SiO40]/[SiO2][H2O]2, as existing during dissolution and at


equilibrium, respectively.

The rate constant, k+, depends not only on pH, I, and T, but also on the ratio of the
interfacial area between the solid phase in contact with the solution and the mass of
solution, on the free energy of formation (Go) of an activated complex
(SiO22H2O*) at the interfacial surface, on the frequency of decomposition of this
activated complex (represented by kBT/h, where kB is Boltzmanns and h is Plancks
constant), and on the activity coefficients of this complex and of silicic acid in the
system (Rimstidt and Barnes, 1980).

Taking this a step further, the time to achieve a certain degree of saturation can be
formulated by integration of (eq.SiI-30) with the following result (Rimstidt and
Barnes, 1980):

t = (1/k)ln{(1Q/K)/(Q0/K)} (eq.SiI-32)

where Q0 = Q at t = 0.

When Q0 = 0, that is starting with pure water, eq.SiI-30 can be recast into

Q /K = 1 ek t (eq.SiI-33)

When a time constant tc = 1/k is introduced, one obtains for t = tc the familiar result

Q/K = (1 e1) = 0.63 (eq.SiI-34)

The reaction can be considered to have gone to completion when t 5tc (0.7%
removed from equilibrium concentration).

18
Silica I integral text 2/22/07

Because eq.SiI-34 only depends on the rate constant for precipitation, k, this
expression holds for all silica polymorphs (Fig. SiI-8).

Fig. SiI-8: time constant (tc) vs temperature for three different of exposed
surface area and volume of water, A/M. The A/M ratios are dimensionless
quantities, calculated according to (A/M) = (A/M)/(A0/M0), where A/M is
exposed surface area (m2) per kg water and A0 1 m2 and M0 1 kg.
(redrawn from Rimstidt and Barnes, 1980).

It is obvious that, apart from the degree of silica saturation and the temperature, the
value of (A/M = surface area per unit of mass) or the geometry of the porous system,
plays a crucial role in the rate of reaction of silica. It can be shown, however, that in
the case of silica precipitation in porous media, the rate of change in surface thickness
or the growth of silica on pore walls, rsfc, is independent of (A/M) and only dependent
on the degree of saturation (17)

rsfc = Vqtz k+ (1 Q/K) (eq.SiI-35)

19
Silica I integral text 2/22/07

where Vqtz is the molar volume of quartz.

VI. Thermodynamic properties


The differences in solubility and kinetics of silica dissolution under various conditions
at or near the Earths surface are reflected in thermodynamic properties of the silica
polymorphs. In the context of this chapter, the prominent silica polymorphs are:
amorphous silica, cristobalite and quartz. The following thermodynamic data cover a
temperature range up to 327C (Table SiI-2).

-Quartz
T (K) G0f H0f S0 T Cp
kJ mol1 kJ mol1 J mol1 J mol1
273 37.658 41.819
298 856.281 910.700 41.439 44.509
400 837.649 910.864 55.842 53.429
500 819.375 910.543 68.469 59.649
600 801.198 909.894 79.778 64.382

Cristobalite (CT)
T (K) G0f H0f S0 T Cp
kJ mol1 kJ mol-1 J mol1 J mol1
273 43.363 44.885
298 854.019 907.864 43.363 44.885
400 835.587 908.006 57.829 53.601
500 817.516 907.633 70.570 60.723
525 813.014 907.466 73.574 62.400
Temperature
transition
(-CT-CT)
525 813.014 906.150 76.080 62.250
600 799.737 905.720 84.300 62.775

Amorphous silica
T (K) G0f H0f S0 T Cp
kJ mol1 kJ mol1 J mol1 J mol1
273 44.456 41.465
298 849.233 901.554 48.475 44.061
400 831.308 901.793 62.666 52.392
500 813.702 901.603 75.003 58.071
600 796.161 901.141 85.972 62.170

Table SiI-2: Thermodynamic data for in the T-range from 0C to 327C for quartz,
cristobalite, and amorphous silica (Navrotsky, 1994).

Obviously, -quartz is the most stable phase (lowest Gibbs energy) under all
conditions, so occurrence of the other phases has a kinetic rather than an energetic

20
Silica I integral text 2/22/07

cause. Since the structure of amorphous silica is not defined as exact as the structure
of the other phases, the properties of natural amorphous silica (opal) may vary.

The T-dependent equilibrium constants for dissolution of silica are (Rimstidt and
Barnes, 1980):

Quartz logK = 1.881 2.028 103 T 1560 T1 (eq.SiI-36a)


-CT logK = 0.0321 988.2 T1 (eq.SiI-36b)
-CT logK = 0.2560 793.6 T1 (eq.SiI-36c)
Amorphous silica logK = 0.3380 7.889 10 T 840.1 T (eq.SiI-36d)
4 1

More recently, slightly different solubility equations are suggested for quartz and
amorphous silica in the temperature range of 8 to 310C (Gunnarson and Arnrsson,
2000):

Quartz logK = 34.188 + 197.47 T1 5. 851 106 T2 +


+ 12.245 logT (eq.SiI- 36e)
Amorphous silica logK = 8.476 485.24 T 2.268 10 T +
1 6 2

+ 3.068 logT (eq.SiI-36f)

The T-dependent dissociation constants of silicic acid (H4SiO40 H3SiO41 + H+) are
(Henley, 1984):

T (K) 273 298 400 500 573


pK1 10.28 9.82 8.97 8.89 9.22

These values pertain to systems with saturated vapor pressure. It has to be recalled
that the dissociation constant of H2O also varies with temperature therefore pH is T-
dependent as well.

The pressure dependence of the free energy of dissolution is dependent on the partial
molal volume change of the reaction

SiO2(s) + 2 H2O Si(OH)4(aq) (eq. SiI-1)

The following pertinent partial volume (changes) in seawater (t = 02C) are given (in
cm3 mol1) (Willey, 1982):

V1 = 9.9, VSi(OH)4(aq ) = 55 5 , VH2O(l) = 18.0 , VSiO 2(s ) = 27.9 (dehydrated) to 35.5 (SiO2
with 9.3% H2O).

These data are useful for the marine system, because temperatures are low and
pressures high at the floor of the deep oceans.

For freshwater systems, VSi(OH)4(aq ) = 60 2 cm3 mol1 (Applin, 1987). Usuallylake


Baikal, for instance, being an exceptionthe pressures remain low enough to neglect
pressure corrections on the equilibrium constant.

21
Silica I integral text 2/22/07

For modeling porewater profiles of dissolved silicon, the following diffusion


coefficients (in units of 105 cm2 s1) can be adopted (Applin, 1987):

D0 H 4 SiO 04 = 2.2 and D0 H 6Si 2 O 07 = 1.0

VII. Hydrothermal systems


To satisfy scientific curiosity as well as for exploration and exploitation purposes, it is
of interest to get a realistic estimate of the subsurface temperature of hydrothermal
systems. Among several others, the silica thermometer can serve this purpose.
Assuming that solid silica phases control the dissolved silica concentrations, the
following relations between these concentrations and subsurface temperatures have
been proposed for the temperature range from 0 to 250C (Table SiI-3).

T (K)
Quartz (no steam loss) 1309/{5.19log (SiO2)}
Quartz (maximum steam loss) 1522/{5.75log (SiO2)}
Chalcedony * 1032/{4.69log (SiO2)}
-Cristobalite 1000/{4.78log (SiO2)}
Amorphous silica 731/{4.52log (SiO2)}

Table SiI-3: Formulas for the calculation of dissolved silica concentrations and
subsurface temperatures from 0C to 250C (Truesdell, 1984). In this
table, the dissolved silica concentrations are in mg kg1. * Chalcedony is
a fibrous variety of quartz.

A graphical representation of the dependency of the solubility of silica polymorphs on


temperature is given in figure SiI-9.

Other geothermometers are: Na/K, Na-K-Ca, K/Mg, Na/Li. The concentrations of


these cations in the hydrothermal system are supposed to be in equilibrium with other
common or not so common mineral phases (Truesdell, 1984). Convergence of the
results obtained with these thermometers gives the most realistic temperature
estimates, but often the results lie quite apart. When this is the case, other means of
reasoning have to come into play to get the most realistic estimate.

Field observations as well as experimental studies show that in active hydrothermal


systems with quickly rising hot water, very little silica precipitates underground as
long as water temperatures are < 230C, whereas above this temperature level quartz,
chalcedony or amorphous silica will precipitate to such extent that the silica
thermometer would indicate too low underground temperatures. As long as no silica
precipitates, the ascending water will react with minerals in the wallrock and,
provided that equilibrium is reached with wallrock minerals, other geothermometers
will be of use. When, however, silica does precipitate in the channels of fluid
advection, attainment of equilibrium with the wallrock minerals is shut off and the
silica thermometer as well as other geothermometers will give faulty results. The best
chance to get realistic temperatures with the silica thermometer is when the fluid can
ascend rapidly through large fractures (A/M will be low). When the fluids have to pass

22
Silica I integral text 2/22/07

through fine-grained sediments (high A/M), the fluids may re-equilibrate at each point
along the flow path and mixing with surface waters becomes more likely. Examples
of the influence of the ascension rate and of the A/M ratio on the deviation of the
silica-thermometer-derived T and the in situ T are given in figures 10a and 10b.
(Rimstidt, and Barnes,1980).

Fig. SiI-9: Solubilities of various silica polymorphs in water at the vapor pressure of
the solution (redrawn from Fournier, 1985)

Fig.SiI-10: a) The effect of the ascension rate (107 to 103 m/s) of a hydrothermal
solution on the quartz geothermometer temperature, assuming a
geothermal gradient of 138C km-1 and an A/M ratio = 100; b) The effect
of the A/M ratio (101 to 104) on the quartz geothermometer temperature,

23
Silica I integral text 2/22/07

assuming a geothermal gradient of 138C km-1 and an ascension rate of


104 m/s. Redrawn from Rimstidt and Barnes (1980).

Most of the silica precipitation occurs in the t-path from 300 to 200C, because
cooling of the fluids during ascension makes them oversaturated. However, continued
cooling will also lower the rate constant of precipitation, k+. Therefore, the decrease in
k+ will gradually surpass the effect of increasing supersaturation. For systems with
underground temperatures < 200C, the rate of precipitation is negligible and, apart
from the effect of possible mixing with surface waters, the dissolved silicon
concentrations can be used as reliable geothermometer. Conversely, when the
underground temperature is > 200C, the silica thermometer will give results that tend
to cluster around 200C (Rimstidt and Barnes, 1980).

VIII. Natural concentrations in the hydrosphere


Chemical weathering can be taken as the starting point of the silica cycle. Natural
waters acquire their silicon concentrations from the congruent (complete) or
incongruent (incomplete) dissolution of (alumino)silicate minerals. A detailed
description of these processes is given in the chapter on weathering. Assuming that
carbon dioxidea weak acidplays a major role in chemical weathering, a general
dissolution reaction can be represented by

primary mineral + (x+0.5y)CO2 + nH2O (secondary mineral)


+ x(alkali ions) + y(earth alkaline ions) + (x+0.5y)HCO3 + zH4SiO40

Realizing that acidity generated by oxidation of sulfidic minerals (essentially pyrite)


contributes to chemical weathering as well, another dissolution reaction can be
formulated as:

primary mineral + xFeS2 + 3.75xO2 + nH2O xFeOOH + (other secondary mineral)


+ 2xSO42 + (4xy)(alkali ions) + 0.5y(earth alkaline ions) + zH4SiO40

For unmetamorphosed sedimentary rocks, amorphous silica serves as an additional (in


case of carbonate rocks) or practically sole (in case of evaporites) source of dissolved
silicon.

The liberated silicic acid usually percolates through the soil, enters the shallow
groundwater which can flow into brooks, streamlets, streams and lakes, or reaches the
deep groundwater. Ultimately, all groundwater flows into the sea. In the course of
flow, dissolved silicon may be partially consumed by secondary processes such as
sorption, reverse weathering, and biological uptake. An estimate of the net discharge
of dissolved Si by rivers into the sea is 5 Tmol y1 (Trguer et al., 1995).

The rate of chemical weathering depends on climate (annual temperature and


humidity) and related vegetation, on rock types and their exposure, and on relief. A
rough distinction between temperate and tropical zones gives the following results for
the average dissolved Si concentrations in river waters (Meybeck, 1987):

24
Silica I integral text 2/22/07

temperate rivers 140 M, tropical rivers 220 M.

This difference can be mainly attributed to temperature differences. For nonvolcanic


drainage areas, riverine silicon concentrations increase by a factor of around 4 from
arctic to tropical climate zones (Berner and Berner, 1996). This observation is
reflected in the distribution of clay minerals in the various climatic zones; soils have
lost relatively more silicon (e.g. laterites) in tropical than in the temperate to arctic
climates.

On a global scale, the contribution (in %) of various rock types to the total dissolved
Si (~5.4 Tmol y1, equal to Trguer et al.s (1995) 5.4 Tmol y1 gross fluvial Si
export) is: shales (35.1), sandstones (16.6), gneisses and schists (11.7), carbonate
rocks (11.3), volcanic rocks (11.1), granites (10.9), serpentinites, marbles,
amphibolites etc. (1.5), gabbros (0.75), evaporites (0.9) (Meybeck, 1987):

The concentration of silicic acid in rivers and lakes can vary seasonally, due to uptake
by flora (e.g. diatoms) in the growing seasons, and liberation from decaying organic
matter in other seasons (Berner and Berner,1996). Average freshwater lakes in
temperate climates have dissolved silica concentrations in the range of 30 to 170 M.
In deep lakesprone to stratificationthe concentrations of dissolved silica can drop
considerably during spring and summer, whereas the hypolimnion becomes enriched
in silicic acid due to dissolution of biogenic silica settling to the lake floor. In arid
climates, lakes have generally higher silicon concentrations (up to ~1.2 mM) and soda
lakeswith their extremely high pHs (even >10)can have concentrations of several
mmol L1.

Shallow groundwaters are usually low in dissolved silicon, in the order of < 100 to
x100 M, depending on the petrographic nature of the underground increasing to
values as high as 1.7 mM in deep groundwaters. Concentrations in hot springs depend
on depth and concomitant temperature of their source, but may be modified by
underground precipitation of silica or dilution with meteoric water; silicic acid
concentrations range from ~ 1 to 10 mM (Wedepohl, 19., .)
Ocean surface waters have usually very low silicic acid concentrations (often < 15
M), because silicon is an essential element for growth of diatoms, silicoflagellates,
sponges and radiolaria. After their deaths, the tests of these creatures settle to the sea
floor and are subject to dissolution during their descent and at or even below the
sedimentwater interface. Seawater is undersaturated with respect to silica; only
within the porewaters of sediments the dissolved silica concentrations may reach their
saturation values. More about these processes is discussed in the next chapter.

IX. Controls on silicon concentrations


a) Role of vegetation
In the analysis of the Si cycle, almost all emphasis is put on the weathering of
(alumino)silicate minerals (next section) as the means of liberation of Si and its
fluvial transport to the seas/oceans, the uptake by diatoms and other siliceous

25
Silica I integral text 2/22/07

organisms, and the subsequent burial in the seafloor and ocean floor. The cycle is
closed by tectonic subduction, volcanism, denudation and(or) uplift (chapter SiII).
Not important on geological time scale of the Si cycle of steady-state equilibrium
between Si liberation on the continents and burial in marine sediments, but relevant
on the shorter time scale is the temporal storage of Si on land. The Si fixation of
phytolith6 silica on the continents (60 200 Tmoly1) is on the same order as fixation
of Si by silica organisms in the oceans (~ 240 Tmoly1). So, on the shorter time scale,
phytoliths are an important part of the Si cycle through the terrestrial bioshpere.
Following the summary in Conley (2002), the points to be stressed are: 1) Si is an
essential element for growth and health of plants; 2) Si is absorbed as H4SiO40 and
deposited as amorphous silica in specialized silica cells of grass, reed7, hairs,
protective spines, leaves, stems, and roots. Si is a compression-resistant structural
component of cell walls; 3) the fraction of Si in terrestrial plants ranges from 110%
or even higher on dry matter basis. As a global average, 13% of total dry weight of
all primary production by vegetation is Si, the molar Si/C ranges from 0.0120.040,
whereas in marine diatoms this ratio is ~0.15.; 4). For deciduous forests, the Si
fixation rates are 930 320 molha1y1 or 93 32 mmolm2y1 (biomass 640 350
mmolm2), for coniferous forests: 280 140 molha1y1 or 28 14 mmolm2y1
(biomass 320 175 mmolm2); 5) forest clearance increases dissolved Si in soil
water, groundwater and streams; 6) building of dams results in trapping of Si with an
effect on its availability for Si organisms in downstream reaches of rivers, in estuaries
and coastal seas.
Alexandre et al. (1997) reported a similarity of phytoliths and soil organic matter
(SOM) contents in soil profiles in a rain forest in equatorial Congo. About 92% of the
biogenic silica input is rapidly recycled, the remaining 8% supplies a stable pool of
phytoliths with a low turnover. Plant uptake of Si increases the chemical weathering
rate without increasing the denudation rate. Uptake, storage and release of Si by
vegetation have to be taken into account when using dissolved Si for tracing chemical
weathering reactions. In figure SiI-11, an overview of the Si fluxes between the
biological and soil compartments is shown.
Derry et al. (2005) concluded, on the basis of GeSi ratios, that plants take up
considerable quantities of H4SiO40 from soil solutions. This Si is recycled in the soil
plant cycle (cf. Figure ), which is significant compared to chemical weathering and
hydrological output. Most of the H4SiO40 in Hawaiian stream waters has passed
through the biogenic Si pool, whereas watermineral interactions account for a
smaller fraction of the fluvial silica flux.
Meunier et al. (1999) reported the existence of a 15-cm-thick phytolith-rich horizon
on Runion Island. This layer developed from trachytic ashes of an age between 3900
and 245 ka. This layer is the relic of a presence of a bamboo forest. The turnover rates
of Si in this soilplant system were estimated to be 0.055 0.01 mol Sim2y1 which
is in the same range as the presently high weathering rates of silica in the area.
Storage of phytoliths may be significant and may retard the output of Si to rivers and
oceans.

6
Phytoliths are the opaline amorphous silica compounds in plants (Conley, 2002). Another description
(Meunier and Colin, 2001) is: phytoliths are fossil micrometric minerals (generally hydrated opal-A
precipitated in plant tissues).
7
The Si-armored edges of grass and reed leaves are often razor-blade sharp and can cause painful
incised wounds.

26
Silica I integral text 2/22/07

Fig. SiI-11: Cycling of silica through soil and vegetation in rain forest in equatorial
Congo. The Si inventories in mol m2 are shown in regular fonts, the Si
fluxes in mol m2 y1 in italicized fonts (after Alexandre et al., 1997). In
steady state, the export of silica in the short time frame is through
groundwater, in the longer time frame through denudation.

Lucas (2001) resumed the results obtained by himself and several other researchers on
the role of Si in vegetation. The Si in plants ranges from 1>10% in dry matter. Most
values in litterfall are 48 mg.g1 dry matter. The resistance of phytoliths to
dissolution varies (in grass land > in forests). Comparison of Si dynamics in a soil
vegetation system is summarized in table SiI-4

Cycling Soil type Turnover rate Phytolith storage Output to


groundwater
mmol m2y1 molm2 mmolm2y1
high Amazonian ferrasols 160 20 61 70 30
high Vosges,France 100 30 1.1 0.7 ~40
medium Cambi- or alfisols
medium Amazonian podzols ~75 ? ~16
low Vosges, France ~16 ~4.6 ~75
podzols

Table SiI-4: Silicon dynamics in some soilvegetation systems in the Amazon and
Vosges (E. France) regions (Lucas, 2001).

Gustafson et al. (2001) reported equilibrium Si concentrations of 0.2 mM at 8C and


0.35 mM at 25C in a Bs horizon. Farmer et al. (2005) noticed that reported
equilibrium Si concentrations all lie in the range 0.080.4 mM, typical for phyloliths.

27
Silica I integral text 2/22/07

They concluded that Si in soil solution reached an equilibrium concentration in the


range of phylolith solubility. In combination with other indicators, such as T-
dependence of solubility, and rate of equilibration, this indicated that phyloliths are
plausible candidates for the principal soured of soluble silica in Bs horizons. The
principal carriers of phyloliths in litter fall are leaves and needles.
Bartoli and Souchier (1978) investigated the typical Si contents (in percent of oven-
dry material): beech, 0.21.1; birch, 0.140.19;, fir, 0.380.66; spruce, 0.141.75;
pine, 0.0250.12. Also reported was that the Si contents increase with age of the
foliage (Hhne, 1963). In concert with Derry et al. (2005), Farmer et al. (2005) come
to the conclusion that the main source of Si for plant uptake by plants are soil
phyloliths, although the ultimate source of Si is mineral weathering.
Fraysse et al. (2006) studies the solubility and dissolution kinetics of bamboo
phytoliths from Runion Island. Acidbase properties are represented by pHIEP8 = 1.2
0.1 (as is) or 2.5 0.2 (organic matter removed). The pKsp0 = 2.74, equal to the
value for vitreous silica. The dissolution rate dependence of pH is represented by

R = k1{> Si(OH) +2 }n + k 2{> SiOH 0 } + k 3{> SiO }m (eq. SiI-37)

where {>i} represents the concentration of surface species (in the order of protonated,
neutral, deprotonated), n and m represent the order of proton- and hydroxyl-promoted
dissolution rates. The minimum dissolution rate was observed for pH 3.
Using results of surface titrations in a surface complexation model, the following rate
constants were, by trial and error, found for the dissolution rate R (in mol cm2 s1)
and {>i} (in mol m2):

k1 = (2.0 0.2) 108 s1 (n = 1)


k2 = (3.13 0.3) 1012 s1
k3 = (0.03 0.01) m2mol1s1 (m = 2.0 0.2)

The upshot of this section is that dissolved Si concentrations in soil water,


groundwater and streams, depend on parameters such as recycling of Si in the soil
vegetation system, seasonal effects, human use of topsoil, storage of phytoliths, tree
cutting, etc.. Only when there is a good reason to assume steady state in the soil
vegetation system, dissolved silica may be considered to be representative of chemical
weathering of primary minerals. In many cases this assumption will be unwarranted
but a detailed lack of the full understanding of the short and long term Si balances
leaves no room for more sophistication.

b) Chemical weathering

Apart from the control of silicic acid concentrations exerted by silica polymorphs,
secondary minerals may play a role in regulating these concentrations. In the general
weathering reactions given above, silicic acid was liberated from primary minerals,
unstable under the conditions at the earths surface. Depending on the type of
secondary minerals formed, part of the liberated silicon can also be consumed in the
formation of secondary minerals such as clays. Assuming attainment of equilibrium
between the silicic acid concentration and the secondary minerals, stability diagrams

8
IEP is the iso-electric point, which is the pH at which the surface charge changes from
positive to negative by adsorption or of protons.

28
Silica I integral text 2/22/07

can be used to visualize the equilibrium relations. Most commonly, these diagrams are
presented as a function of the concentrations of dissolved silicon and a specific cation
and of the pH. Some basic reactions are (Faure, 1998):

2 albite + 9 H2O + 2 H+ kaolinite + 2 Na+ + 4 H4SiO40


log [Na+]/[H+] = 2 log [H4SiO40] 0.19 (boundary 1)

kaolinite + 5 H2O 2 gibbsite + 2 H4SiO40


log[H4SiO40] = 4.68 (boundary 2)

3 albite + 12 H2O + 2 H+ paragonite + 2 Na+ + 6 H4SiO40


log [Na+]/[H+] = 3 log [H4SiO40] 4.1 (boundary 3)

2 paragonite + 3 H2O + 2 H+ 3 kaolinite + 2 Na+


log [Na+]/[H+] = + 7.63 (boundary 4)

paragonite + 9 H2O + H+ 3 gibbsite + Na+ + 3 H4SiO40


log [Na+]/[H+] = 3 log [H4SiO40] 6.4 (boundary 5)

2 albite + 4 H2O + 2 H+ pyrophyllite + 2 Na+ + 2 H4SiO40


log [Na+]/[H+] = - log [H4SiO40] + 2.48 (boundary 6)

pyrophyllite + 5 H2O kaolinite + 2 H4SiO40


log [H4SiO40] = 2.67 (boundary 7)

silica(am) + 2 H2O H4SiO40


log [H4SiO40] = 2.7 (boundary 8)

(albite: NaAlSi3O8; gibbsite: Al(OH)3; kaolinite: Al2Si2O5(OH)4;


paragonite: NaAl3Si3O10(OH)2; pyrophyllite: Al2Si4O10(OH)2).

These relations can by represented in a stability diagram (Fig. SiI-12) in which the
cation/hydroniumion activity ratio is plotted against the silicic acid activity.

Similar equilibrium equations can be used for other primary (e.g. potassium feldspar,
anorthite) and secondary (e.g. illite, montmorillonite, zeolites) minerals. The problem
with some of these minerals is that they have no fixed composition, so phase
boundaries depend on the composition of the mineral phase(s) present. A schematic
stability diagram of such phase relations is given in figure SiI-13.

29
Silica I integral text 2/22/07

Fig. SiI-12: Stability diagram (25C and 1 atm) for albite and the weathering
products pyrophyllite (Py), amorphous silica (Si), paragonite (Pa),
kaolinite (K) and gibbsite (G) as a function of activities of silicic acid
and sodium/hydronium ion activity ratio. The numbers at the boundaries
between mineral phases correspond with the boundary conditions
mentioned in the text (redrawn from Faure, 1998).

30
Silica I integral text 2/22/07

Fig. SiI-13: Schematic stability diagram for primary silicates and clay minerals as a
function of silicic acid and cation/hydronium ion activity ratio,
Mn+)/(H+)n. Redrawn from Hershey and Millero (1986).

In a weathering scenario for the earth surface, the initial solution composition starts
off with the chemistry of precipitation (very low if any dissolved silicon, pH slightly
acidic, low cation concentrations), far removed from the stability of the primary
phases. It has to be realized that in these 2D-diagrams a third axis for aluminum
activities is missing, implying that the actual view is at cross sections of a 3D-stability
diagram in which dissolved aluminum is assumed to be constant (and low). For most
systems, with clay minerals forming from primary minerals by percolating solutions,
aluminum is assumed to behave conservatively (no losses); taking into consideration
the stability of aluminum oxide, this is only permitted within a narrow pH-range of
about 5 to 8. The pathway of the solution composition leads from the lower left corner
in the stability diagram (gibbsite formation) towards the upper right corner, but at
phase boundaries silicon will first become incorporated in the next secondary mineral
till all of the former secondary mineral has been consumed in the reaction with excess
silicic from the dissolving primary mineral. In a multicomponent system, with a
combination of primary minerals present in various quantities and dissolving at

31
Silica I integral text 2/22/07

different rates, the silicic acid concentration will be controlled by the assemblage of
secondary mineral phases formed.
The stability diagrams also show that primary phases can be reconstituted by reverse
weathering, provided that silicic acid and cation activities as well as pH become such
that they lie within the stability field of that particular primary phase. Often these
reactions will occur for heterogeneous nucleation at pre-existing surfaces having the
best suited lattice structure. Examples are overgrowths of quartz grains from silica-
supersaturated solutions and of feldspars on detrital ones (Kastner and Siever, 1979).
Sorption of silicic acid occurs in the presence of clay minerals, which can be
considered as an initial step in neoformation of more stable mineral phases from clay
minerals (Fig. SiI-14)

Fig. SiI-14: Solubility of silicon in contact with suspensions of 5 g L-1 kaolinite or


montmorillonite in seawater. The bottom curves represent the solubility
development in time without, the top curves with prior addition (0.262 mM) of
silicic acid (redrawn after Mackenzie et al., 1967).

32
Silica I integral text 2/22/07

X. Lithification and silification


Dissolved silicon concentrations higher than equilibrium concentrations can persist
for quite some time, but ultimately decrease by sorption or precipitation of silica.
Sorbed silicon may serve as a precursor for further sorption and homogeneous
nucleation of silica.
Lithification by cementation can occur in situations of silica-supersaturation of waters
in and at the surface. In the subsurface, overgrowths of silica on quartz grains may
occur with a loss of porosity. Extensive circulation of oversaturated water through the
subsurface is a prerequisite for effective cementation. At or near the surface silica
cementation may form silcrete rocks, late precipitates of chert associated with
weathering surfaces. Host rocks are often argillaceous sandstones; typical examples
are present in Australia in peneplained regions having sporadic rainfall. The
observation is that quartz grains are often partially replaced and surrounded by
microcrystalline quartz, fibrous chalcedony and minor opal to constitute an
interlocked quartzitic mass (Dapples, 1979). Beds of silcrete may be 2 to 4 or 5 meter
thick and extend for many kilometers. The processes leading to the formation of
silcretes are: precipitation of silica from upwardly moving dissolved-silica-rich
solutions by evaporation or by contact with downwardly moving interstitial pore
waters; the climatic conditions favorable of silcrete formation are: alternating wet and
dry seasons in (sub)tropical settings, or arid (sub)tropical settings (Friedman and
Sanders, 1978).

An interesting case of silification is the petrification of hard parts of invertebrates and


of plant material, which often perfectly preserves the original structure of the so
fossilized materials. In this respect, silification is superior over calcification. Well-
known examples of silification are found in the Devonian Rhynie Chert in Scotland
and in petrified wood in Yellowstone National Park (Siever and Scott, 1963).
Woody plants consist of cellulose (45-51%), hemicellulose (23-38%) and lignin (19-
33%), the ratios depending on the type of tree. Chemically, cellulose is a high-
molecular-weight (HMW) glucose polymer, hemicellulose a medium-molecular-
weight polysaccharide (hexoses and pentoses), and lignin a HMW polymer with
aromatic groups. Cell walls in woody tissues have a thin primary wall and a thick
secondary wall with usually three layers. Silification commences by silica deposition
upon the more accessible cellular surfaces, particularly on the inner cell walls
encapsuling the lumina and on the lining of pit chambers, and continues by outward
deposition, filling the voids, particularly the lumina. Upon progressive silification, the
woody templates begin to deteriorate, making space for more silicic acid
incorporation, with the final result that the interior cell walls are filled with silica. The
organic matter gradually disappears by degradation and maceration, but some of it is
often retained in silicified wood (Leo and Barghoorn, 1976).
From a chemical perspective, celluloses, lignins and their degradation products
contain abundant functional, particularly hydroxyl, groups, capable of forming
hydrogen bonds. The process in which vascular tissue serves a a template for silica
deposition can be represented as

33
Silica I integral text 2/22/07

H
:
ROH + HOSi(OH)3 RO OSi(OH)3
:
H

Continuous build-up of these surface complexes triggers interaction and


polymerization with formation of siloxane bonds and elimination of water.
Since petrification is a slow process, the decay of the wood has to be slow as well.
Protection against oxygen is secured in sediments which allow water logging of the
wood, which maintains the wood in a swollen, plastic state and permeable for
exchange with solutions in which the wood is immersed. Sediments with volcanic ash
provide ideal conditions for silification, because intense chemical weathering of basic
ashes liberates large amounts of dissolved silicon. A relatively high pH accompanying
weathering of such material helps to attain high silica concentrations. On the other
end of the pH spectrum, pyrite oxidation in former marine sediments can also raise
the silicic acid concentration and promote silification of organic matter in acid sulfate
soils. Within the wood, however, the pH of permeating fluids during silica
emplacement, has probably remained near neutral or slightly acidic, as strong alkaline
or acidic conditions would have hydrolyzed the organic material and destroyed its
structure (Leo and Barghoorn, 1976).

XI. References
Alexandre, A., Meunier, J.-D., Colin, F., Koud, J.M. (1997). Plant impact on the
biogeochemical cycle of silicon and related weathering processes. Geochim.
Cosmochim. Acta., 61: 677682.

Applin, K. (1987). The diffusion of dissolved silica in dilute aqueous solution.


Geochim. Cosmochim. Acta, 51: 21472151.

Bartoli, F., Souchier, B. (1978). Cycle et rle du silicium dorigine vgtable dans les
cosystmes forestiers tempres. Ann. Sci., For., 35: 187202.

Berner, E.K., Berner, R.A. (1996). Global environment: Water, air, and geochemical
cycles. Prentice Hall, Upper Saddle River, N.J., U.S.A.. 376 pp.

Brady, P.V., Walther, J.V. (1990). Kinetics of quartz dissolution at low temperatures.
Chem. Geol., 82: 253264.

Carroll,S.A., Maxwell, R.S., Bourcier, W., Martin, S., Hulsey, S. (2002). Evaluation
of silicawater surface chemistry using NMR spectroscopy. Geochim.
Cosmochim. Acta, 66: 913926.

Conley, D.J. (1998). An interlaboratory comparison for the measurement of biogenic


silica in sediments. Mar. Chem., 63: 3948.

Conley, D.J. (2002). Terrestrial ecosystems and the global biogeochemical silica
cycle. Global Biogeochemical Cycles, 16: 68-168-8.

34
Silica I integral text 2/22/07

Dapples, E.C. (1979). Silica as an agent in diagenesis. In: G. Larsen, G.V.


Chilingar (Eds). Diagenesis in sediments and sedimentary rocks.
Developments in Sedimentology, V. 25a. Elsevier, Amsterdam, Oxford, New
York. 99141.

Derry, L.A., Kurtz, A.C., Ziegler, K., Chadwick, O.A. (2005). Biological control of
terrestrial silica cycling and export fluxes to watersheds. Nature, 433: 728731
(plus supplementary information).

Dietzel, M. (2000). Dissolution of silicates and the stability of polycyclic acid.


Geochim. Cosmochim. Acta, 64: 32753281.

Dove, P.M., Nix, C.J. (1997). The influence of the alkaline earth cation, magnesium,
calcium, and barium on the dissolution kinetics of quartz. Geochim.
Cosmochim Acta, 61: 33293340.

Dove, P.M. (1999). The dissolution kinetics of quartz in aqeous mixed cation
solutions. Geochim. Cosmochim. Acta, 63: 37153727.

Eggiman, D.W., Manheim, F.T., Betzer, P.R. (1980). Dissolution and analysis of
amorphous silica in marine sediments. J. Sed. Petrol., 50: 215225.

Farmer, V.C., Delbos, E., Miller, J.D. (2005). The role of phytolith formation and
dissolution in controlling concentrations of silica in soil solutions and streams.
Geoderma, 127: 7179.

Faure, G., 1998. Principles and applications of Geochemistry. Prentice Hall, Upper
Saddle River, New Jersey. 600 pp.

Fournier, R.O., Rowe, J.J. (1977). The solubility of amorphous silica in water at high
temperatures and high pressures. Am. Mineralogist, 62: 10521056.
\
Fournier, R.O. (1985). The behavior of silica in hydrothermal solutions. In Berger,
B.B., Bethke, P.M. (eds). Geology and Geochemistry of Epithermal Solutions.
Reviews in Economic Geology, Volume 2. The Economic Geology
Publishing Company, El Paso, Texas, U.S.A., 4561.

Fraysse, F., Pokrovsky, O.S., Schott, J., Meunier, J.-D. (2006). Surface properties,
solubility and dissolution kinetics of bamboo phytoliths. Geochim.
Cosmochim. Acta, 70: 19391951.

Friedman, G.M. , Sanders, J.E. (1978). Principles of Sedimentology. John Wiley


& Sons, New York etc.. 792 pp.

Grasshoff, K., Ehrhardt, M., Kremling, K. (1983). Methods of seawater analysis.


Verlag Chemie GmbH, Weinheim, Germany. 419 pp.

Gunnarssson, I., Arnrsson. S. (2000). Amorphous silica solubility and the


thermodynamic properties of H4SiO40 in the range of 0 to 350C at Psat.
Geochim. Cosmochim. Acta, 64: 22952308.

35
Silica I integral text 2/22/07

Gustafson, J.P., Berggren, D., Simonsson, M., Zysset, M., Mulder, J. (2001).
Aluminium solubility mechanisms in moderately acid Bs horizons of
podzolized soils. Eur. J. Soil. Sci., 52: 655665.

Henley, R.W. (1984) . pH Calculation for hydrothermal fluids. In: R.W. Henley,
A.H. Truesdell, P.B. Barton Jr, J.A. Whitney. Fluid-mineral equilibria in
hydrothermal systems. Reviews in Economic Geology, Volume 1. The
Economic Geology Publishing Company, El Paso, Texas, U.S.A., 8398.

Hershey, J.P., Millero, F.J. (1986). The dependence of the acidity constants of silicic
acid on NaCl concentration using Pitzers equations. Mar. Chem., 18: 101
105.

Hhne, H. (1963). Analysis of leaves in younger spruce plantations. Arch. Forstwes.,


12: 341 360.

Icopini, G.A., Brantley, S.L., Heany, P.J. (2005). Kinetics of silica oligomerization
and nanocolloid formation as a function of pH and ionic strength at 25C.
Geochim. Cosmochim. Acta, 69: 293304.

Iler, R.K. (1973). Effect of adsorbed alumina on the solubility of amorphous


silica in water. J. Colloid Interface Sci., 43: 399408.

Iler, R.K. (1979). The chemistry of silica. John Wiley & Sons, New York etc., 866 pp.

Isshiki, K., Sohrin, Y., Nakayama, E. (1991). Form of dissolved silicon in seawater.
Mar.Chem., 32: 18.

Kaim, W., Schwederski, B. (1994). Bioinorganic chemistry: Inorganic elements


in the chemistry of life. John Wiley & Sons, Chichester etc., 401 pp.

Kastner, M., Siever, R. (1979). Low temperature feldspars in sedimentary rocks.


Am.J.Sci., 279: 435479.

Koning, E., Brummer, G-J., Van Raaphorst, W., Vn Bennekom, J., Helder, W., Van
Iperen, J. (1997). Settling, dissolution and burial of biogenic silica in the
sediments off Somalia (northwestern Indian Ocean). Deep-Sea Res. II, 44:
13411360.

Langmuir, D. (1997). Aqueous environmental geochemistry. Prentice Hall, Upper


Saddle River, New Jersey, 600 pp.

Leinen, M. (1985). Techniques for determining opal in deep-sea sediments: A


comparison of radiolarian counts and X-ray diffraction data. Mar.
Micropaleont., 9: 375383.

Leo, R.F., Barghoorn, E.S. (1976). Silification of wood. Bot. Mus. Leaflets
(Harvard University), 25: 146.

36
Silica I integral text 2/22/07

Lucas, Y. (2001). The role of plants in controlling rates and products of weathering:
Importance of biological pumping. Ann. Rev. Earth Planet. Sci., 29: 135163.

Mackenzie, F.T., Garrels, R.M., Bricker, O.P., Bickley, F. (1967). Silica in


seawater: Control by silicate minerals. Science, 155: 14041405.

Marshall, W.L., Warakomski, J.M. (1980). Amorphous silica solubilities II.


Effect of aqueous salt solutions at 25 oC. Geochim. Cosmochim. Acta, 44:
915924.

Mazer, J.J. , Walther, J.V. (1994). Dissolution kinetics of silica glass as a function of
pH between 40 and 85C. J. Non-Crystalline Solids, 170: 3245.

Meunier, J.D., Colin, F., Alarcon, C. (1999). Biogenic silica storage in soils. Geology,
27: 835838.

Meunier, J.D., Colin, F. (2001). Phytoliths: Applications in Earth Sciences and


Human History. A.A. Balkema Publishers, Lisse/Abingdon/Exton
(PA)/Tokyo. 378 pp.

Meybeck, M. (1987). Global chemical weathering of surficial rocks estimated


from river dissolved loads. Am. J. Sci., 287: 401428.

Mortlock, R.A., Froelich, P.N. (1989). A simple method for the rapid determination of
biogenic opal in pelagic marine sediments. Deep-Sea Res., 36: 14151426.

Mller, P.J., Schneider, R. (1993). An automated leaching method for the


determination of opal in sediments and particulate matter. Deep-Sea Res., 40:
425444.

Navrotsky, A. (1994). Thermochemistry of crystalline and amorphous silica. In:


P.J. Heany, C.T. Prewitt and G.V. Gibbs (eds). Silica: Physical Behavior,
Geochemistry and Materials Applications. Reviews in Mineralogy, Volume 29
Mineralogical Society of America, Washington D.C.. 309329.

Perry, C. C. (1999). Biogenic Silica: A model of Amorphous Structure Control.


Chapter 11 in: B. Janveit and P. Meakin (eds). Growth, Dissolution
and pattern Formation in Geosystems. Kluwer Academic Publishers,
Netherlands. pp. 237 251.

Rimstidt, J.D., Barnes, H.L. (1980). The kinetics of silica-water reactions.


Geochim. Cosmochim. Acta, 44: 16831699.

Schlter, M., Rickert, D. (1998). Effect of pH on the measurement of biogenic silica.


Mar. Chem., 63: 8192.

Siever, R., Scott, R.A. (1963). Organic geochemistry of silica. In: I.A. Breger.
Organic geochemistry. Pergamon Press, London. 579595.

37
Silica I integral text 2/22/07

Spadini, L., Schindler, P.W., Sjberg, S. (2005). On the stability of the AlOSi(OH)32+
complex in aqueous solution. Aquatic Geochem., 11: 2131.

Stumm, W. (1992). Chemistry of the SolidWater Interface. John Wiley & Sons, New
York etc., 428 pp.

Tallberg, P., Hartekainen, H., Kaireselo, T. (1997). Why is soluble silicon in


Interstitial and lake water samples immobilized by freezing? Water Res., 31:
130134.

Tossell, J.A. (2005). Theoretical study on the dimerization of Si(OH)4 in aqueous


solution and its dependence on temperature and dielectric constant. Geochim.
Cosmochim. Acta., 69: 283292.

Trguer, P., Nelson, D.M., Van Bennekom, A.J., DeMaster, D.J., Leynaert, A.
and Quguiner, B. (1995). The silica balance in the world ocean: A reestimate.
Science, 268: 375379.

Truesdell, A.H. (1984). Chemical geothermometers for geothermal exploration.


In: R.W. Henley, A.H. Truesdell, P.B. Barton Jr, J.A. Whitney. Fluid-mineral
equilibria in hydrothermal systems. Reviews in Economic Geology, Vol. 1.
The Economic Geology Publishing Company, El Paso, Texas, U.S.A. 3143.

Van Cappellen, P., Dixit, S., Gallinari, M. (2002). Biogenic silica dissolution and the
marine Si cycle: kinetics, surface chemistry and preservation. Ocanis, 28:
417454.

Walther, J.V. (1996). Relation between rates of aluminosilicate mineral dissolution,


pH, temperature, and surface charge. Am.J.Sci., 296: 693728.

Wedepohl, K. H. (19691978) Handbook of Geochemistry . Springer, Berlin

Weres, O., Yee, A., Tsao, L. (1981). Kinetics of silica polymerization. J. Colloid
Interface Sci., 84: 379402.

Willey, J.D. (1982). Partial molal volume calculations for the dissolution of aged
amorphous silica in salt water and seawater at 02C. Geochim. Cosmochim.
Acta, 46: 1307 1310.

Wirth, G.S., Gieskes, J.M. (1979). The initial kinetics of dissolution of vitreous
silica in aqueous media. J. Colloid Interface Sci., 68: 492500.

Wollast, R., Chou, L. (1985). Kinetic study of the dissolution of albite with a
continuous flow-trough fluidized bed reactor. In: J.I. Drever (ed). The
Chemistry of Weathering, Reidel, Dordrecht. 7596.

Xiao, Y., Lasaga, A.C. (1994). Ab initio quantum mechanical studies of the
kinetics and mechanisms of silicate dissolution: H+ (H3O+) catalysis. Geochim.
Cosmochim. Acta, 58: 53795400.

38
Silica I integral text 2/22/07

Xiao, Y., Lasaga, A.C. (1996). Ab initio quantum mechanical studies of the
kinetics and mechanisms of quartz dissolution: OH catalysis. Geochim.
Cosmochim. Acta, 60: 22832295.

39

Você também pode gostar