Sunteți pe pagina 1din 33

CHAPTER 3

Polymerization of Silica
From the time of Graham ( I ) , who made an intensive study of sols and gels, many attempts have been made to explain the behavior of silicic acid. When freshly made by acidifying a soluble silicate or hydrolyzing the ester, silicic acid is not colloidal, since i t diffuses easily through parchment or animal membranes and has a molecular weight by freezing point depression corresponding to monomer. Soon the molecular units become larger and pass through membranes only slowly and then not at all (2). This could be because the monomer or other small primary particles form aggregates, or because the individual particles increase in size and decrease in number. Freundlich appeared to recognize these alternatives when he wrote: Whether it is rather a matter of polysilicic acids, which give larger micellae, being formed from simple silicic acid, or whether the crystalloid particles originally present already consist of polysilicic acids, but are exceedingly fine amicrons which continually increase in size-cannot yet be said with certainty. In his terminology, a micella is a colloidal particle in which foreign substances (ions, water) are present in its structure, that is, a porous aggregate, whereas the amicron is a discrete parlick too small to be seen with the ultramicroscope. He recognized that such particles in a colloidal solution could consist of one very large molecule, in other words, a single unit, not an aggregate. Because the most obvious behavior of a silicic acid solution is that it increases in viscosity and finally forms a gel, its polymerization was generally assumed to be an aggregation process or a polymerization by which smaller molecular units linked together into larger ones. The nucleation and growth of discrete particles prior t o the stage where aggregation begins have not been recognized by many workers, who held to the idea that Si(OH), polymerized into siloxane chains which then were branched and cross-linked as in many organic polymers. Even now attempts are still made to apply the idea of monomer functionality and condensation polymerization theory of organic chemistry t o the silica system. In fact, there is n o relation or analogy between silicic acid polymerized in an aqueous system and condensationtype organic polymers. [However, the behavior of silica in molten glasses is quite a
I72

Polymerization of Silica

173

different matter, and in that system the conventional polymer theory has been shown by Masson (3) to be applicable.] In 1925, Kruyt and Postma (4a) pointed out that there are two groups of silicic acid sols. The first group has a p H of 4.5 or less, and the viscosity of the sol increases with time. On the other hand, pure silica sols, having a p H of 7 or higher, a r e relatively stable, the viscosity either remaining the same or decreasing with time. This difference in behavior is explained as follows. The more alkaline sols bear a negative charge and are thereby stabilized. However. the addition of soluble salts lowers the charge of the particles and causes gelation or flocculation. On the acid side, where there is essentially no charge, aggregation or flocculation occurs, causing an increase in viscosity, and eventually gelation. Tourky (4b) also discussed the structural differences between silicic acids in acidic and basic solutions; in acidic solutions, fibrillar o r network structures arise through the formation of oxygen bridges between silicic acid units. It was Carmen (4c) who first clearly stated that silicic acid polymerizes to discrete particles which then aggregate into chains and networks. The formation of silica gel can be regarded as taking place in two stages. In the first, initially formed Si(OH), condenses to form colloidal particles. In dilute solution, a further slow increase in particle size is the only subsequent change, but at a concentration of about I percent silica, these primary particles are able to condense together to give a very open but continuous structure, extending throughout the medium, thus bestowing a certain degree*ofrigidity upon it. In both stages of polymerization, the mechanism is the same, that is, condensation to form Si-0-Si links, but in the first stage, condensation leads to particles of massive silica, while in the second, since it is not possible to fit two particles accurately together over a common face, the number of Si-0-Si linkages between particles is fewer in number than those within the particles themselves. They are merely sufficient to bind adjacent particles together, in a fixed position relative to one another, and thereby lead to a rigid, highly porous, tangled network of branching chains. . . . Thus three stages are actually recognized:
1. Polymerization of monomer to form particles.

2. Growth of particles. 3. Linking of particles together into branched chains, then networks, finally extending throughout the liquid medium, thickening it to a gel.
Since Carmen published this in 1940, further experimental data continue to confirm his point of view. There is general agreement that polymerization, that is, the reactions that result in an increase in molecular weight of the silica, involves the condensation of silanol groups: -SOH

+ HOSi-

-SiOSi-

+ H,O

T h e term polymerization is used in its broadest sense, the mutual condensation of S i ( 0 H ) t o give molecularly coherent units of increasing size, whether these are

174

Polymerization of Silica

spherical particles of increasing diameter or aggregates of an increasing number of constituent particles. Formation and growth of spherical particles is one kind of polymerization that takes place under certain conditions. Aggregation of particles to form viscous sols and gels is another kind of polymerization occurring under other conditions. Both types of polymerization may occur at once.

GENERAL THEORY OF POLYMERIZATION


The general theory of polymerization is first outlined. Then the details of each step are reviewed and finally the more recent work of a number of investigators is discussed. Succeeding steps in polymerization from monomer to large particles and gels or powders have been represented schematically by ller ( 5 ) as in Figure 3.1. This applies to aqueous systems, in which silica is somewhat soluble. Very little is known about the polymerization when Si(OH), is formed in nonaqueous solutions. The individual steps are as follows. Later each step is considered in detail in the light of individual investigations. (a) Monosilicic acid is soluble and stable in water at 25'C for long periods of time i f the concentration is less than about 100 ppm as SiO,. When a solution of
MONOMER

DIMER

CYCLIC

Figure 3.1. Polymerization behavior 0 1 silica. I n basic solution ( B ) particles in sol grow in sire w i t h decrease in numbers: in acid solution or in presence of flocculating salts ( A ) , particles aggregate into three-dimensional networks and form gels.

General Theory of Polymerization

175

monomer, Si(OH),, is formed at a concentration greater than about 100-200 ppm as S O 2 , that is, greater than the solubility of the solid phase of amorphous silica, and in the absence of solid phase on which the soluble silica might be deposited, then the monomer polymerizes by condensation to form dimer and higher molecular weight species of silicic acid. (b) The condensation polymerization involves an ionic mechanism. Above pH 2 the rate is proportional to the concentration of O H - ion and below 2 to the H + ion. (c) Silicic acid has a strong tendency to polymerize in such a way that in the polymer there is a maximum of siloxane (Si-0-Si) bonds and a minimum of uncondensed SiOH groups. Thus at the earliest stage of polymerization, condensation quickly leads to ring structures, for example, the cyclic tetramer, foll'owed by addition of monomer to these and linking together of the cyclic polymers to larger three-dimensional molecules. These condense internally to the most compact state with SiOH groups remaining on the outside. (d) The resulting spherical units are, in effect, the nuclei that develop into larger particles. The solubility of these very small particles depends on the particle size, that is, the radius of curvature of the surface. It also depends on the completeness of the dehydration of the internal solid phase. If the latter is formed at ordinary temperature it may contain uncondensed O H groups but if formed above 80"C, and especially above pH 7, it is almost anhydrous. (e) Because small particles are more soluble than larger ones (Chapter I ) and since not all the small three-dimensional particles are the same size, the particles grow in average size and diminish in numbers as the smaller ones dissolve and the silica is deposited upon the larger ones (Ostwald ripening). However, the higher solubility of smaller particles is pronounced only when the particle size is smaller than about 5 nm and very pronounced when it is less than 3 nm. Hence above pH 7, where the rate of dissolution and deposition of silica is high, particle growth continues at ordinary temperature until the particles are 5-10 nm in diameter, after which growth is slow. However, at low pH, where the rate of polymerization and depolymerization is slower, particle growth becomes negligible after a size of 2-4 nm is reached. At higher temperatures, growth continues to larger sizes, especially above pH I . The very early formation of particles was also proposed by Vysotskii et al. ( 6 ) , who studied the early stages of polymerization and similarly stated that there are two basic processes of particle growth of silica in the aqueous system:
I . Growth of particles at the expense of silicic acid in the solution from the moment

of its preparation. 2. Further growth of larger particles by deposition of silicic acid dissolving from the smaller particles. This is a slower process and may be negligible at low pH after the monomer has been used up.
(r) Above pH 6 or 7, and up to 10.5, where silica begins to dissolve as silicate, the silica particles are negatively charged and repel each other. Therefore they do not collide, so that particle growth continues without aggregation. However, if salt is

I76

Polymerization of Silica

present at a concentration greater than 0.2-0.3 N , as when sodium silicate is neutralized with acid, the charge repulsion is reduced and aggregation and gelling occur. It is paradoxical that under some conditions, precipitation or gelling is prevented by raising the temperature. I n this pH range a sol of 2-370 silica with a borderline salt concentration of 0.2-0.3 N gels i f aged at ordinary temperature, However, if the sol is first heated to 80-IOOC the particles grow in size and decrease in number so that aggregation and gelling are greatly retarded or even prevented permanently. (g) At low pH the silica particles bear very little ionic charge and thus can collide and aggregate into chains and then gel networks. If the concentration of SiO, is more than I % such aggregation may begin as soon as the first small particles are formed. However, at lower concentrations and at pH around 2, the monomer is converted largely to discrete particles before they begin to aggregate. On the other hand, at pH 5-6, monomer is converted rapidly to particles which simultaneously aggregate and gel so that i t is not possible to separate two processes. The rate of aggregation increases rapidly with concentration so that in any case above 1 % silica, aggregation probably involves not only particles but also oligomers. The process of aggregation and gelling in the silica system is unique because, unlike other metal oxides, the solid phase remains completely amorphous and appreciably soluble in water and is generally in solubility equilibrium with the monomer. It is essential to understand that while sol is being converted to gel, the growing aggregates contain the same concentration ojsilica and water as in the surrounding sol regions. These aggregates or "gel phase" cannot be seen because the density and refractive index of the gel phase are the same as those of the remaining sol. Thus before the sol solidifies only a slow increase in viscosity can be noted, with little change in other properties, u p to a point where the viscosity begins to increase rapidly and solidification occurs at the "gel point." The most common way of determining the "gel point" is to observe when the meniscus of a sol in a container no longer remains horizontal when the container is tilted. It may be difficult to visualize how particles in a suspension can rearrange themselves into three-dimensional networks without changing the silica concentration. I n certain microscopic regions in the sol the particles arrange themselves in chains, and these in turn branch and form networks. These can be isolated in sols gelling at low pH by adding an inert miscible fluid such as water or alcohol. With a twofold dilution, all the particles not attached to a three-dimensional network move apart. but the rigid networks retain their structure and thus are more dense than the medium, and so can be separated by centrifuging. In this way the percentage of silica that has been converted to "gel phase" can be measured. Gelling occurs when about half of the silica has entered the gel phase, which can be thought of as spherical solidified regions in suspension which cause a rapid increase in viscosity when the "volume fraction" reaches about 0.5. After the gel network has been formed, the structure becomes stronger as the necks between particles become thicker owing to solution and deposition of silica. I n [his chapier, attention is concenirated on the details of the various processes o j polynierization up to the point 0.f gel formation. The completion of gel formation and subsequent changes in structure are dealt wirh in Chapter 6.

Monosilicic Acid
Overall Effect of pH on Gelling

177

Whether polysilicic acids or larger particles of colloidal silica are involved, the general effects of pH are generally as indicated schematically in Figure 3.2. Curve A B C represents the behavior of silica in the absence of salts. Sols have a maximum temporary stability with longest gel time around p H 1.5-3, and a minimum stability with rapid gelling around p H 5-6. Above about p H 7, no gel is formed since the particles are charged and only particle growth occurs. Curve DEF represents the general behavior when an electrolyte such as NaCl or Na,SO, is present a t a concentration above about 0.2-0.3 N . The salt lowers the ionic charge on particles. At low p H both sols gel, and salt has little effect. In the neutral region the pH of minimum stability is higher ( E ) when salt is present.

MONOSILICIC ACID Monomeric silicic acid, Si(OH), has never been isolated. It is a very weak acid and exists only in dilute aqueous solution, since it polymerizes when it is concentrated. I t

I
2

I
4

1
6
PH

I
8

I
IO

The effect of pH on the gelling of silica sols. Curves A-C-sols in the absence o f sodium salts; D-F: in the presence of sodium salts.
Figure 3.2.

178

Polymerization of Silica

is the soluble form of silica that is in equilibrium with the solid phases. As described in Chapter I , it is a neutral highly hydrophilic, essentially nonionized substance that cannot be isolated from water. In pure form, if it could be prevented from polymerizing, it might be expected to be a clear liquid resembling glycerin. The basis for this speculation is that a very low molecular weight polysilicic acid was isolated as a clear viscous anhydrous liquid by Robinson (7), who found that it polymerized immediately to clear, hard silica gel when exposed to a trace of atmospheric moisture. It was highly hygroscopic, and soluble in polar organic solvents such as alcohol, but insoluble i n hydrocarbons. In view of its unique character, a brief description of its preparation is justified. To a violently agitated mixture of 83.2 grams tetraethylene glycol dimethyl ether, 52 ml H,O, and 40 ml20% H2S04,was added, in a thin stream, 125 ml of a solution of sodium silicate containing 122 g I - ' SiO, and 38 g I-' Na,O. After 10 min, 50 g anhydrous Na,SO, was added and the mixture stirred, then allowed to stand. The supernatant clear liquid layer of the ether contained silicic acid weighing 186 grams. This was separated and at once subjected to vacuum distillation at ordinary temperature to remove the water. Then the high-boiling ether was extracted with an equal volume of benzene, leaving a viscous, water-clear, anhydrous polysilicic acid. Sufficient H,SO, was present to make the liquid acidic, which gave it temporary stability. It soon set to a clear hard gel, especially when exposed to moisture or warmed. Monomer can be removed from solution by strong-base ion-exchange resin, presumably because it is ionized to HSi0,- by the O H - ions at the resin surface, and is then adsorbed.

Preparation
The various methods of preparing monosilicic acid may be summarized as follows. A saturated solution of monosilicic acid, Si(OH),, containing about 0.01% SiO,, is obtained when pure amorphous silica is equilibrated with water at room temperature. A more concentrated (supersaturated) solution can be obtained only indirectly by liberating monosilicic acid from its compounds under carefully controlled conditions; at low temperature and low pH, dilute solutions remain supersaturated with respect to amorphous silica for appreciable periods. For example, at pH 3 and OC, solutions of monosilicic acid up to 0.1 M (0.6% SiO,) can be prepared by spontaneous hydrolysis of monomeric silicon compounds, sich as silicon tetrachloride or methyl orthosilicate, and also by reacting monomeric silicates, such as sodium or magnesium orthosilicates or hydrated crystalline sodium metasilicate, with dilute acid.
Dissolving Silica

Jander and Heukeshoven (8) reported that amorphous silica gel gives a true solution of silica in water; later Jander and Jahr (9) found that the silica in solution had a diffusion coefficient of 0.53, indicating a molecular size about equivalent to Si(OH),.

Monosilicic Acid

I79

Subsequently Alexander, Heston, and Iler (IO) verified that amorphous silica exhibits a definite equilibrium solubility in water, amounting to about 0.0 1-0.012% SiO, in the saturated solution. Kitahara and Oshimo ( 1 1, 12) dissolved quartz under high pressure a t 400C to obtain a solution containing 350-400 ppm which was quenched to obtain a solution of monomer at pH 6 which then polymerized. A 400 ppm solution of monomer was obtained by saturating water with silica gel at 95-IOOC (13). Egorova (14) reports that i n the pH range 1.2-3.7 it remains unpolymerized even after 2 hr. This appears to be a convenient way to make and store a 400 ppm solution of monomer for polymerization or deposition studies a t 25"C, where the solubility of gel is around 100-120 pprn.
Hydrolysis of Monomeric Silicon Compounds

Dilute solutions of monomer can be obtained by hydrolysis of halides, esters, or acyl derivatives such as silicon tetraacetate; SiCI, requires the later removal of HCI. Thus Willstatter, Kraut, and Lobinger (15) led SiCI, vapor into water a t 0C while adding silver oxide to maintain p H 3 and t o precipitate the chloride. This method was also investigated by Gruner and Elod (16). The very rapid hydrolysis of tetraacetate gave monosilicic acid, whereas ( C H , S 0 0 ) , S i 0 S i ( C H , C 0 0 ) , gave disilicic acid according to Schott and Fischer (17). Using a cryoscopic method, they demonstrated that in this system a t 25C Si(OH), was most stable a t p H 2.8 and the disilicic acid a t 3.1. The relative rates of polymerization of monomer and dimer, each at its most stable pH, were compared a t unspecified concentrations. The monomer was said to be more stable than the dimer but this probably depends on the p H and the method of preparation. Ethyl orthosilicate has been used to prepare silicic acid that is not completely monomeric because the two-phase hydrolysis is not instantaneous. Disilicic acid was prepared from tetraethyl orthosilicate by the Brintzingers (18). Methyl orthosilicate hydrolyzes rapidly to monomer. Brintzinger and Troemer ( 1 9) obtained essentially monosilicic acid by hydrolyzing methyl orthosilicate in 0.001-0.01 N l HCI. Weitz, Franck, and Schuchard (20) prepared monosilicic acid by hydrolyzing tetramethyl silicate in 0.002 NI HCI at room temperature, checking the molecular weight by the freezing-point method. The methyl ester gives monomer when hydrolyzed in about IO sec in water at p H 3. Thus monosilicic acid can be obtained as a 0.13 M solution (0.8% SO,) by hydrolyzing methyl orthosilicate in N H,SO, or HCI solution a t 25"C, and it remains practically unchanged for 2-3 hr as described by Schwarz and Knauff (21). They hydrolyzed the ester in an apparatus from which the methanol and some water could be vacuum distilled and the molecular weight of the silicic acid was determined by the freezing-point method. After 24 hr, the molecular weight corresponded t o that of dimer, but there was no indication that it was more stable a t this point. At silica concentrations around 0.5-1% SO,, it is impossible to preserve the monomeric state at 25"C, even a t the pH of optimum stability. Five minutes after the preparation of a 0.083 M solution of Si(OH), (0.5% S O ) , ) a t p H 2. some disilicic acid is present (22).

180 Dissolving Mononieric Silicates in Acid

Polymerization of Silica

Monomeric crystalline silicates dissolve and are neutralized to liberate monosilicic acid at about pH 2. Kraut (23) prepared monosilicic acid by dissolving sodium metasilicate hexahydrate i n various acidic solutions at low temperature. He reported that monosilicic acid is most stable at around pH 2-3. Weitz, Franck, and Schuchard (20) demonstrated that when Na,SiO,. 9H,O was reacted with acetic acid it liberated Si(OH),. Also olivine (magnesium orthosilicate, Mg,SiO,) dissolved in 1.0 N HCI to give a practically 100% yield of monosilicic acid, the solution containing 0.04% SiO,. Thus monosilicic acid may be liberated from silicates which contain Si0,lions separated by cations. such as are present in anhydrous orthosilicates. Alexander (24a) found that sodium metasilicate hydrolyzed to disilicate when dissolved in water unless NaOH is added to form orthosilicate, Na,SiO,. However, if crystalline Na,SiO,. 9H,O and strong-acid H ion-exchange resin are added N solution of H,SO, at OC to maintain pH 3, a 0.1 M simultaneously to a solution of monomer can be obtained. Thilo, Wieker, and Stadt (24b) made monomer by dissolving Na,SiO, glass in cold dilute acid. With glass of composition 1 SO,: 1.5 Na,O, disilicic acid was obtained. Contrary to Schott and Fischer (17), they found disilicic acid is much more stable than the monomer and can be made reproducibly at higher concentrations. Coudurier, Baudru, and Donnet (25) modified Alexanders method by maintaining the pH at 2.5, and Okkerse (29) preferred 0.01 N HCI at pH 2 as the reaction medium. Silicic acid was prepared by Funk (26) by dissolving certain mineral orthosilicates (monosilicates in German) in a solution of HCI in anhydrous methanol. The calcium, barium, and magnesium chlorides were soluble in methanol and the resulting monosilicic acid was much more stable than in water. Soluble minerals were beta and gamma Ca,SiO, (dicalcium silicate); Ca,OH(HOSiO,) (dicalcium silicate alpha hydrate); Ca, ( H O S i 0 3 ) , . 2 H , 0 (synthetic aufwillite); BaOH(H,SiO,) .4H,O; Mg,SiO, (synthetic forsterite); and Mg,AI,(SiO,), (garnet). The monosilicic acid polymerizes and is precipitated when water is added. I t would seem likely that it is present at least partially as a methyl ester. Funk and Frydrych (27) prepared solutions in which 90% of the dissolved silica was Si(OH), at concentra:ions up to I % , by dissolving anhydrous Ca,SiO, in dry methanolic HCI. Polymerization was very rapid unless the solution was diluted at once to 0. I % S O , . By adding acetone, the CaCI, was precipitated, leaving a relatively pure solution of silicic acid (28).

Characteristics of Silicic Acid Since Si(OH), has never been isolated or even obtained in a concentrated solution without considerable polymerization, very little is known about its physical or chemical properties. Most measurements have therefore been made in very dilute solutions.

Monosilicic Acid
Diffusion Conslant

181

This was measured in seawater by Wollast and Garrels (30) and found to be 1.0 *
0.05 x IO-5 cmz sec-'.

Ionization Conslants
The ionization constant of monosilicic acid has been evaluated in many ways. As already discussed in Chapters I and 2, the pK, appears to be about 9.8 at 25C determined by Roller and Ervin (31) in a system involving calcium oxide, silica, and water. More recently, careful measurements by Marsh, Klein, and Vermeulen (32) of the equilibrium between Si(OH), and HSi(OH),- over a range of pH led to a value of pK, = 9.9. Measurements on a system containing extremely pure silica by Schwartz and Muller (33) give a still more precise value of pK, = 9.91 i 0.04. They hydrolyzed extremely pure methyl orthosilicate in water in a system which rigorously excluded atmospheric impurities, to obtain solutions containing from 12.4 to 155 ppm SO,. The conductivity and pH were measured with precision at 25C using low frequency alternating current. From the initial conductivities the value of the acidity constant was calculated. In 0.5 M NaCIO, solution Bilinski and Ingri (34a) found that monosilicic acid, Si(OH),, had a first dissociation constant at 25C corresponding to pK, = 9.46 *
0.02.

The values of the ionization constants according to Scherban (34b) are as follows:
Kl
=

[H'] [H3SiO;] [HSiO,l [ H + ][H,SiO:-] [H3SiO:] [H'] [HSi03-] [H,SiO:-] [H'] [SiO'-] (HSiOS,-]
=

2 x lO-'O

K, K, K,

2 x 10-l2 2 x IO-'*

2 x IO-',

The increase in the first ionization constant of monosilicic acid with temperature was measured by Seward (34c). The data were obtained in the presence of borax buffer at 0.1-0.6 M concentrations. The value of pK, ranged from 8.88 i 0.15 at 130C to 10.0 0.2 at 350C. The ionization behavior of Si(OH), and the formation of polysilicate ions i n 1-5 M NaCl solutions at silica concentrations of 0.005-0.05 m have been measured with precision by Busey and Mesmer (34d). They found negligible formation of any complexes between monomer and the sodium ion to form NaO(OH), in solution. (The complexing behavior of polysilicic acid is of course quite different.)

182

Polymerization of Silica

Increase in Ionization Constant with Polymerization

The ionization constants of disilicic and polysilicic acids, colloidal silicas, and gels are pertinent to the polymerization of monomer and SO are considered here. The increasing acidity of silicic acid upon polymerization was reported by Belyakov et a). (35). As the monomer polymerized, the pK, was determined by titration and the degree of polymerization by the cryoscopic method. The maximum pK, was reported as 10.7 for H,Si,O, but then decreases to 6.5 for high polymers. However, it is not known whether the decrease occurs upon the formation of ring compounds with =Si(OH), groups or when three-dimensional particles with =SiOH groups on their surfaces have been formed. The most doubtful point is the reported pK, of 10.7 for disilicic acid, which would mean it is a weaker acid than monosilicic acid, of which the pK, is 9.8. Analogy with other inorganic acids would suggest that disilicic acid should be a stronger acid than monomer. Unfortunately the dimer is difficult if not impossible to prepare and keep in sufficiently pure state for strength measurements, although a solution in which probably at least 50% of the silica was dimeric was prepared by Coudurier, Baudru, and Donnet (36). Another indication that polysilicic acid is a stronger acid than Si(OH), is furnished by ion-exchange studies by Strazhesko and others (37, 38). The ionization constants of acid centers on the polysilicic acid surface are at least two or three orders of magnitude higher than the constant of monomeric Si(OH),. Dugger et al. (39a) estimated the acidity of the silanol groups on silica by measuring the ion exchange of H + with 20 metal ions. By this means they showed that the first hydrogen to leave the pure silanol surface must have a dissociation constant k, of 10-4-10-8.This is much more acidic than Si(OH),, which has a k, of The variation of pK, of the silica surface with degree of neutralization was found to be as shown in Figure 3.3 by Strazhesko et al. (39b), who carried out studies of the mechanism of ion exchange on silica gel using N a + , C a t , Cs+, Caz+,SP+,and Bal, and also on gel in the form of divalent and trivalent metal salts. Duffy and lngram (39c) have been able to estimate the ionization constants of a wide variety of acids from the electro-negativity of the constituent elements and also from the Lewis basicity or optical basicity from spectroscopic data. Although this has failed when applied to monosilicic acid it might give useful data for polysilicic acids (39d). Allen, Matijevic, and Meites (39e) also developed an equation relating surface change on particles of colloidal silica and pH which indicated that the pK, for the surface entirely in the hydrogen form is 6.4 and completely in the sodium form is 9.6. The acidity of silanol groups on the surface of silica or polysilicic acid has been examined by Schindler and Kamber (40), who calculated the intrinsic acidity constant, K,,,, from results of titrating silica gel at 25C in 0. I M NaCIO, solution:

Monosilicic Acid

183

0.25

0 50

0 75

DEGREE OF NEUTRALIZATION

Figure 3.3 Relation between pK, of silanol groups on the surface of amorphous silica and the degree of neutralization according to Strazhesko (39b).

With regard to the equation

log K

log [H+) + log

1-a

where a is the degree of neutralization or the fraction of silanol groups that are ionized at a given pH. The intrinsic acidity constant K,,, is defined as equal to K when a approaches zero. From the data they developed the following equation:
1.9(a

log K

log K,,, -

+ aZ) 0.039 + u
1 -u
a

where log K i n , = -6.81. Then log [H-] = logK,n, 1.9(a

+ a2) 0.039 + a

log

To obtain this equation it was necessary to take into account the silicate species in solution at equilibrium. The following ionization constants for these species in 0.5 M NaCIO, solution at 25OC were reported by Bilinski and Ingri (34a).

184

Polymerization of Silica

This assumes the solubility of silica gel is I20 ppm.

[H') [(HO),SiOl-] [( HO), SiO-]

10-1258

K:

The total exchange capacity, C, defined as the maximum number of ionizable groups under the conditions of titration, was found to be 2.43 O H groups nm-, in 0.1 N NaCIO, solution. As a matter of interest, the authors report 3.4 O H nm-, in I M NaCIO, and 5.83 in 3 M NaCIO, solution. This approach to the acidity of the silica surface is quite different from that of Yates (41, 5), which was developed as a modification of the conventional equation applicable to organic polyacids: pH
=

pK - k , log,, a N - k , log,,

; t -9

where R is the ratio of the molar concentrations CSIOz/CN.,O (where Na,O is the titratable alkali), A is the specific surface area of silica in square meters per gram, and a and N are the activity and normality of sodium salt in the system. Based on Bolt's data (421, the constants were found to be pK = 12.08, k, = 0.74, k , = 3.47, and k , = 2430. In this formula there is no assumption made as to the fraction of silanol groups that are ionizable, but the number of charges per unit area can be calculated from CNs10 and A . As an example, a point was taken on Schindler and Kamber's curve for silica gel of specific surface area of 372 m z g - ' . At a = 0.1, log K was - 8 . 3 . From their equations a pH of 7.35 was calculated. This is a reference to the "a" of page 183. Then taking the Yates equation with pH = 7.35 and A = 372, a value of R was found to be 240 S O , : Na,O. Thus for 60 x 240 grams of SiO, there are present 2 x 6 x IOzs Na* counterions and ionized O H groups. Then the number of ionized sites per square nanometer is 12 x 1 0 2 3 = 0.22 60 x 240 x 372 x 10" Schindler assumed there are 2.43 ionizable O H groups nm-, so a = 0.22/2.43 = 0.09. This is very close to the value of 0.10 for the point originally selected. In other words, the two approaches appear to lead to similar results.

Monosilicic Acid

185

The acid dissociation constant of the O H groups on polymeric silica has also been shown to be about lo-'.' by an entirely different method. Hair and Hertl (43) measured the frequency shift of the infrared absorption band of phenolic hydroxyl groups when adsorbed on the surface of silica and compared these with the shifts of phenol in the presence of alcohols of known acidity constants. Marshall et al. (44) concluded from a similar study that the pK, of some SiOH groups on the silica surface may be about 7.2.
Isoelectric Point

The isoelectric point (iep) of Si(OH), in solution in the absence of colloid or solid phase has apparently not been measured but presumably it would be between pK, and pKb, where these are the negative logarithms of the equilibrium constants for:

K,: Si(OH),
Kb: Si(OH),

=
=

(HO),SiO(HO),Si+

+ H+

+ OH-

The latter equation might also be written H,O

+ Si(OH), '= (HO),SiOH: + O H -

Most measurements have involved solutions in which both Si(OH), and polymeric silica or colloid or solid phase were present. An exception is the case where the initial rate of polymerization of Si(OH), has been measured at different pH values. Here the initial step is one of the following: Si(OH),

+ -OSi(OH), = (HO),SiOSi(OH), + O H Si(OH), + +Si(OH), = (HO),SiOSi(OH), + H +

Presumably, then, the pH at which monomer reacts most slowly with itself to form dimer might correspond to the iep of Si(OH),. Okkerse (29) measured the rate of disappearance of molybdate-reactive silica from solution and found it to be at a minimum between pH 2 and 3. In a study by Goto (45) on the rate of disappearance of monomer from a solution of 2400 ppm SiO, at 25"C, a minimum at pH 2.0-2.2 was found.
SILICON. The existence of a cationic form of monomeric silica is of CATIONIC course implied in the assumption that Si(OH), has an iep. Colloidal particles of silica have been shown to carry a positive charge at low pH, but direct proof that silicon can exist as a cation has not been available. It is therefore interesting that in very dilute solution (66 ppm) monomeric silica has been shown to react with HCI to form the ion (H,O,Si(OH),+CI- according to Cherkinski and Knyaz'kova (46). This was determined by the difference in precise conductivity measurements of 0.0025 N

186

Polymerization of Silica

N a O H solution titrated with 0.005 N HCI, with and without the presence of 0.001 I M %(OH),. The chloride compound exists only in very dilute solution.
Point ofzero Charge

The point of zero charge (pzc) where the surface charge is zero and the isoelectric point where the electrical mobility of silica particles is zero have been measured by many methods. De Bassetti, Tschadek, and Helmy (47) measured the pzc for silica gel by a calorimetric method, from which they concluded the value must be between 2 . 5 and 3 . However, the d a t a may not preclude a value as low a s p H 2 since the heat of neutralization becomes exceedingly small below p H 3. I n an extensive study of silica polymerization, De Boer, Linsen. and Okkerse (48) found the iep to be between p H 1 and 1.5, and that condensation was slowest there, as shown by several means including viscosity studies. Vysotskii and Strazhesko (49) have pointed out that there has been relatively little attention paid to the pzc or iep of silica in spite of the fact that in other colloid systems they are key factors. These authors recalled the observation o f Freundlich ( 2 ) that whereas lyophobic colloids are least stable at the iep, the lyophilic colloid, silica, appeared t o be the most stable. This is not quite true because colloidal silica is permanently stable when it is negatively charged a t pH 9-10, but there is, as Freundlich recognized, a marked temporary stability maximum at the iep around pH 2 (see Figure 3.2). Vysotskii and Strazhesko show that in the presence of a given acid such as sulfuric, the iep is not only the point of minimum rate of gelling but also of syneresis and is also the point at which gels of maximum strength and maximum specific surface area are obtained. All these characteristics result not only because the rate of aggregation is a t a minimum at the iep, but also because the rate of growth of the ultimate particles from monomer is at a minimum, so that the ultimate particles are smallest as they form the gel. These authors noted the relation between the p H of slowest gelling and the pK, o f the acid used. Their data are plotted in Figures 3.4 and 3.5. Their p H values for HNO,, H,SO,, and H,CrO, are similar to those reported by Iler (50), who also reported a number of other very strong acids which gave maximum gel times at about the same p H as for HCI and HNO,, for example, N H 2 S 0 3 H , HCIO,, and CH(SO,H),. However, the point was not brought out that with weaker acids such as acetic, although the sol may be most stable at p H 3.5, it is nevertheless far less stable than the sols made with stronger acids at p H 1.5-2.0. Similar results were reported by Tai and Kiang (51), hydrochloric, sulfuric, and nitric acid giving a maximum gel time at pH 2 , phosphoric a t 2 . 5 , and acetic a t 4.0. It was proposed that the polymerization rate is proportional to

The iep appears t o be at about pH 1.5, according to ion-exchange studies by Vysotskii and Strazhesko (52) and Kirichenko and Vysotskii (53) of ion-exchange

20

IO

0
0 1 2 3 4

PH
Figure 3.4. Effect of p H on gel time at 25C of silicic acid sols of different concentrations made from H,SO, and sodium silicate: Curve I : 1.09 M , 65.4 gI . I SiO,. Curve 2: 1.33 M , 79.8 g1-I SO2.Curve 3: 1.78 M , 106.8 g1-l SO,. [From Vysotskii and Strazhesko (49).]

+5

-5

I pH

OF

MAXIMUM GEL TIME

Relation betw,een the pH of slowest gelling rate and pK, of the acid used for neutralizing the sodium silicate. [From Vysotskii and Strazbrsko (49).]
Figure 3.5.

I87

188

Polymerization of Silica

sorption of rubidium ions from 0.1 N RbNO, solution on silica gels pretreated a t temperatures up t o 1ooO"C. All curves in Figure 3.6 converge to this p H a t zero adsorption. This general approach is summarized by Klimentova, Kirichenko, and Vysotskii (54). In summary, iep and pzc of silica have been variously reported t o be from p H 0.5 to 3 . 7 according t o a review of the literature on this point by Parks ( 5 5 ) , who cited 12 references. However, a p H of around 2 f 0.5 appeared to be an average for various types of silica ranging from purified ground quartz to colloidal silica. S o m e variation may be expected, depending on whether the surface is crystalline or amorphous, possibly on particle size, and especially on the presence of impurities. The question remains how the iep determined from maximum gel time or minimum rate of disappearance of monomer relates to the polymerization mechanisms involved. The relation between the isoelectric point of polysilicic acid and the stability of sols, rate of gelling, and properties of resulting gels has been summarized by Klimentova, Kirichenko, and Vysotskii (54). This behavior can be summed up by saying that all the phenomena observed involve the formation and hydrolysis of Si-0-Si bonds, and that the rates of these reactions depend on a catalytic effect which is a t a minimum a t p H 1.5-2.0'in the presence of anions of strong acids and the minimum becomes greater at higher pH in the presence of anions of weaker acids. From the fact that the rate of disappearance of monomer by polymerization is second order above p H 2 and third order below 2. Okkerse (29) concluded that an anionic form of silica was involved above p H 2 and a cationic form below 2. Thus the isoelectric point must be at p H 2. Similarly, De Boer, Linsen, and Okkerse (56) considered that the isoelectric point is around p H 2, since the polymerization rate is

? 0
x

e
Y

5:
a W a c
v,
2 2

a 3

90

50

5
W

a 0 a

(z

Y o
1 2 3 4
PH

The isoelectric point. Adsorption of rubidium ions versus pH on silica gels preheated to various temperatures. Curves 1-5. temperatures 300, 500, 700, 900, and IOOO"C, respectively. (From Kirichenko and Vysotskii (53).]
Figure 3.6.

Monosilicic Acid

189

a function of H + and O H - on each side of this point. In further work (57) they found by electrophoresis studies that the iep was between pH 1.0 and 1.5 in a 0.5% SiO, sol and at pH 2 when the sol was diluted to 0.26%. It was also shown that there was a sharp minimum in the viscosity at pH 1.9. Similar observations were made by Tai An-Pang (SS), who related gel time to the ionization constant of silica. The significance of the iep of silica in the silica-water system involving Si(OH), and polymerized or solid silica surfaces is still not clear, but the preponderance of evidence suggests that for monomeric Si(OH), the iep may be between pH 2 and 3, and for polymeric forms between 1.5 and 2.
Stability of Monomeric Silica

As long as the concentration of Si(OH), is below the equilibrium solubility of amorphous silica, usually assumed to be about 120 ppm for silica gel but around 70-80 pprn for vitreous silica, it has been assumed that monomer would remain, as such, in water solution at 25C. However, such a solution is supersaturated with respect to quartz and probably to other crystalline species(Chapter I).There is also a possibility that a solution of monomer at a concentration of 100-150 ppm might nucleate a particular less soluble polymeric species of lower solubility. Such a case may be involved in the observations of Schwartz and Muller (33), who made a highly purified solution of silicic acid from methyl orthosilicate at concentrations up to 150 ppm. Initially, conductivity measurements indicated that the silica was monomeric, but after half an hour the conductivity, at all concentrations, slowly decreased to about half the original value. This happened even though in half the samples the concentrations were less than the solubility of amorphous silica. It was assumed that the monomer polymerized slowly at pH 7 to a polymer species that is smaller than usual colloidal dimensions, since it passed through an ultratiler, yet it must be more insoluble than amorphous silica. Unfortunately, this change was not followed by means of the molybdic acid method to see whether it involved simple dimerization at pH 7, which might have escaped the notice of previous investigators. However, if this were the case, and if disilicic acid has a pK, of - 10.7 as reported by Belyakov et al. (35), then the conductivity would have decreased by much more than 50%. (This is discussed later in further detail.)

Reactions of Monosilicic Acid In view of the relatively neutral character of Si(OH), with its physical resemblance to an organic polyol, it is not surprising that at pH 2, where it is not ionized, few if any interactions with other substances have been observed. Its most obvious reaction is self-polymerization to higher molecular weight polysilicic acids which are more reactive. The interaction of polysilicic acids with other substances is considered later in this chapter. However, there are a few reactions in which Si(OH), may take part.

190

Polymerization of Silica

These are interactions either with other acids to form anhydrides or with a few extremely weakly basic metal cations.

Phosphoric and Boric Acids


Silica has long been known to react with anhydrous H,PO, but the wide variety of possible compounds has not been investigated. The reaction is, in effect, a condensation, with water eliminated. For example, by heating amorphous silica with H,PO, at a molar ratio of 1 : 2 for a week at 8O-18O0C, silicon phosphate is formed. Excess A 10%solution H 3 P 0 , is removed with dioxane and the product is dried at 10ODC. can be made in water, giving a 2 . 7 % concentration of silica (59a). Silicon phosphate has long been known but this example of a water-soluble material is mentioned because i t probably hydrolyzes to Si(OH),. The reactions of boric acid with silica appear to parallel those of phosphoric acid since in dilute solution there appears to be no interaction between the acids, but on dehydration at high temperature, Si-0-B bonds are formed in the resulting mixedoxide glass. The Si-0-P and Si-0-B bonds are hydrolyzed in aqueous solution.

Sulfuric Acid
The issuance of a series of patents involving silicon salts of sulfuric acid is surprising since it is unexpected that a reaction product of two of the oldest known chemicals should have escaped attention for so long. However, the existence of silicon phosphate suggests that the sulfate might also exist. Blount (59b)has disclosed the compound silicodihydrogen sulfate, SiO(HSO,), which was obtained by dehydrating dihydroxy silicon dihydrogen sulfate, (HO),Si(HSO,),, with concentrated sulfuric acid. It is claimed that these solid compounds are obtained by stirring powdered Na,SiO,. 5 H 2 0 for several hours in an excess of concentrated H,SO,. Finally the sulfate salt is hydrolyzed i n water giving a white granular silico-formic acid or monosilanol. HSi(O)OH, and monosilandiol, H,Si(OH),(?). However, no further information about properties or analysis is given. I f the products exhibited a characteristic x-ray diffraction pattern or other identifying features their existence as compounds would be less equivocal. I f a crystalline character is retained the compounds might be clathrates with H,SO, within the lattice or exist as a different crystal structure, as in the case of the phosphates. On the other hand, if the powders are amorphous then the) may be microporous silica gels with pores filled with acid; if anhydrous, internal surface groups of = S i O S 0 3 H may be present.

Iron and Uranium


Monomeric silica does not react with most metal ions in water at low pH where Si(OH), can exist, since for reaction to occur it is probable that some hydrolysis to a basic metal ion must first take place.

H,O t Fe3+zFez+OH + H

Monosilicic Acid

191

However, very few metal ions form basic ions at the pH of 2, where monomeric Si(OH), is most stable. Iron and uranium are the only ones which have so far been reported. Monomeric silica reacts with uranyl ion as follows, according to Porter and Weber (60): UO:*

+ %(OH),

UO,SiO(OH);

+ H+

The equilibrium constant for monomer concentrations in the range 0.024-0.035 M is 0.01 0.00 1. As evidence of some chemical combination, certain forms of natural hydrated silica gels and also laboratory-prepared gels impregnated with uranyl salts have been observed by Iler to fluoresce with a strong greenish yellow color under ultraviolet light. The other known reaction of monomer with a metal cation is the case offerric iron, reported by Weber and Stumm (61) and further examined by Porter and Weber in regard to the effect of the degree of polymerization of silica. They polymerized the silica at a concentration of 2280 ppm at pH 9-10 for various lengths of time, conditions that are known to give very small spherical particles. With increasing polymerization of silica with formation of adjacent SiOH groups that can combine with iron, at pH 2, the number of SiOH groups combined per iron ion increases from one on the monomer to two or three as the particles become larger, the radius of curvature larger, and the SiOH groups closer together. The following equation suggested by the authors does not indicate the degree of polymerization of silica, but only the number of SiOH groups that can react with Fe3+,liberating the corresponding number of H ions:
+

Following the absorption characteristics of the iron as it complexes, the following values were obtained: Calculated Diameter (nm)

Mol. Wt.
60 13000 26000
120000
Qn

D.P.
1 217 434 2000

n
1.02 1.67 1.67 1.76

-log[Qn(Si)l
2.76 4.22 4.26 4.48

2.6 3.4 5.6

= =

D.P.

equilibrium constant degree of polymerization

In the stock solutions containing 2280 ppm of silica at pH 9-10 there must have been an appreciable concentration of monomer in equilibrium with the polymer. Based on the calculated particle sizes this would amount to at least 2.6, 2.2, and 1.7

192

Polymerization of Silica

m M or 156, 132, and 102 ppm as monomer, which undoubtedly also combined with the iron but was not taken into account. The interaction of Si(OH)r with ferric iron is evidenced by the fact that concentrations of 10-4-10-3M SiO, in water catalyze the oxidation of Fez+ to Fe3+.Schenk (62) has derived a quantitative relation between the rate of oxidation and the concentration of monomeric silica. Below pH 3.5 a soluble complex between Si(OH), and Fe3+exists. At pH 6-8, a ratio of 3 Si(OH), to 1 Fe3+ prevents precipitation of Fe(OH),. However, in the case of AI3+, a fivefold excess of Si(OH), is required to prevent precipitation.

Chrom iu m

It is peculiar that in view of the similarities of AI3+ and Cr3+ in their precipitation behavior as hydrous oxides these elements are widely different in their interaction with silica. One reason is that the chromite ion is not formed as easily as the aluminate ion, AI0,-. The Cr5+ ion is much larger than AI3+and cannot fit into the SiO, lattice to give stable anions like (SiAIO,)-. T h u s the Cr3+ ions show a peculiar inertness relative to monomeric silica, in marked contrast to the behavior of AISy.When amorphous silica was heated under pressure with a mixture of Cr(OH), and AI(OH), for 2 days at 30O0C, only the alumina combined with silica (63). This behavior of chromium probably explains the rarity of chromium silicate minerals. Hexavalent chromium as H,CrO, appears to form a complex with Si(OH),, according to Her (50). The chromate ion is unique among inorganic anions in that it retards the polymerization of Si(OH), in the pH range from about 0.5 to 3.0. At pH 1.7, where Si(OH), is most stable, the increase in gel time of a 1 M silica sol was linear with Cr03:Si0, ratio. I t ranged from 69 hr with no CrO, to 270 hr at a CrO,:SiO? ratio of 0.75, and at higher ratios was then constant at 270 hr. The latter is the gel time of a 0.5 M SiO, sol in the absence of H,CrO,. Thus the system behaved as though HzCrO, dimerized the silica quantitatively and the excess had no further effect: 0 -H - -0- - H -0 HO-Si-0-Cr-0-Si-OH

0- H - - 0 - -H-0

I I

II II

I I

It appears that this dimer then gelled at a rate as though the concentration of silica was only half the original. Unfortunately, the gel obtained was not examined to see i f the CrO,Z- ion was actually bound in the structure. No compound such as silicon chromate has been reported, but evidently silicon can be linked through oxygen to hexavalent chromium. A chromic acid ester of a silanol group was made by Schmidt and Schmidbaur (64), who prepared the tri-

Monosilicic Acid

193

methylsilyl ester:

0 (CH,),SiOCrOSi(CH,), 0
Aluminum

As discussed in other chapters relating to the effect of aluminum ions, there is a peculiar affinity between the oxides of aluminum and silicon. At this point only a few observations regarding the interaction with monomeric Si(OH), are noted. Aluminum oxide is far less soluble than silica in water at 25"C, p H 5-8, as evidenced by early data by Okura, Goto, and Murai (65), shown in Figure 3.7. Monomeric silica reacts with AI3+ ions and is precipitated most effectively at pH 9, according to G o t o (66). Thus with a solution containing initially 35 ppm monomeric SiO, a t p H 9, the addition of 20-100 ppm of AI as AI3+ ions reduced the silica concentration to a value C, such that A C = 300, where A and C are ppm of AI3+ and SiOz, respectively. However, this probably did not represent true equilibrium. Over a long period of time monomeric silica, (SIOH),, reacts with AIS+ ion at 25C to form colloidal aluminum silicate of the halloysite composition: 2 Si(OH),

+ 2 AI9+ + H,O

A1,Siz0,(OH)4

+ 6 H+

By reacting soluble silica and alumina a t various pH values for periods up to 4 years and measuring the concentrations of residual Si(OH), and AI3+, Hem et al. (67)

IO

i f 1
05
d
J

02
01

6
PH

Figure 3.7. Solubility of aluminum oxide in water versus pH [From Okura. Goto. and Murai

(65)l.

194

Polymerization o f Silica

measured the following constants:

[AI(OH);]' [Si(OH),j2 [H-]'

''

The standard free energy of the colloidal aluminum silicate was -897 i I kcal m ole- I . Monomeric silica is strongly adsorbed onto the surface of hydrous aluminum oxides. There is a reaction between Si(OH), and crystalline AI(OH), by which several reaction layers of SiO, are built up, with simultaneous decrease in p H of the suspension (68a). Formation of the first layer is rapid, but the second and third layers form progressively much more slowly. It would seem that diffusion of AIS+or AI0,- from the surface of the crystal must be involved, with the formation of a silica-rich aluminosilicate. A relatively low content of aluminum ion in the S O , layer greatly reduces its solubility, thus explaining the deposition of SiO, from a solution unsaturated with respect to pure amorphous silica. Baumann (68b) found that when different amounts of aluminum ion were added to a solution of monomer (420 ppm SiO,), more silica remained in the molybdate reactive state than when no aluminum was present. With no aluminum present, after 4 days there remained 130 ppm of molybdate-reactive silica a s monomer in equilibrium with 290 ppm of relatively inactive high polymer. But when aluminum was present in the AI:Si atomic ratio of I : 7 , there remained about 200 ppm of molybdate-reactive silica. I t can be interpreted that the alumina had combined with silica to form an aluminosilicate that later was decomposed by the strongly acidic molybdate reagent liberating additional active silica that appeared as monomer. However, when the silica concentration was only 60 ppm, and thus below the solubility of amorphous silica, no polymerization occurs excepi when alumina is added. I n this case when the AI:Si ratio is I : I to 1 : IO the aluminum ion brings together monomer to form a silica-rich complex in which some of the silica is also linked together into a state that is later less molybdate-reactive. Baumann's extensive data deserve detailed stud). The final reaction product at the alumina surface is halloysite. When a dilute solution of monomeric silica is brought in contact with g a m m a alumina, it is adsorbed at a rate strongly dependent on pH and area A . In a medium of constant ionic strength (0.1 N NaCI) and at silica concentrations of 10-3-10-' M , Huang (69a) found the initial adsorption to be rapid. When below pH 9 the rate is proportional to A [SiO,]' [H+J-O5 , whereas above p H 9 it is proportional to A z [Si0,]'.5 [H '1, Huang proposed that HSiO; is the major reacting species. I t is possible that with a quaternary ammonium base and in the absence of metal cations, aluminosilicate anions may remain in solution, for example, (HO)3SiOAl(OH)zOSi(OH)31~, Flanigen (69b) reported that quaternary ammonium silicate and aluminate remained in solution until a sodium salt was added. I n the case of pure alpha alumina, the writer has found that there is no interaction with monomeric silica. Colloidal alpha alumina free from other forms of alumina or

Characterization of Silicic Acids

195

AI3+ ions has been prepared by treating the particles with 24% H F solution for 24 hr to remove all other types of alumina and silica impurities, and washing with water, then NH,OH t o remove all F - ions from the alumina surface (70a). At p H 7-8, monomeric silica is not adsorbed, nor does it react with this form of alumina even though the specific surface area is 24 m 2 g-. Presumably AI3+or polybasic AI ions are required for reaction with Si(OH),.

Divalent Cations
It is known that the ligand properties of deprotonated Si(OH), with a polyvalent metal cation can lead to a stable complex such as [FeOSi(OH)J3+, as reported by Weber and Stumm (61); however, much less is known about complexes of divalent cations. Santschi and Schindler (70b) measured the stability of complexes involving C a Z + and MEz+ a t around p H 8-9 in 1 M sodium perchlorate solution a t 25C. Complex formation was weak and occurs only in the presence of excess salts. In natural waters, such complexes are not formed.

CHARACTERIZATION OF SILICIC ACIDS


It is not possible to discuss all the techniques used for measuring or characterizing silicic and polysilicic acids and small colloidal particles, but some of the methods, especially applicable t o following the polymerization, are reviewed.

Reaction with Molybdic Acid


The history and use of this reaction in analyzing for silica is discussed in detail in Chapter 1 and its application in characterizing silicate ions in Chapter 2. Further refinements and use of this indispensable reaction for studying the polymerization of silica are now described. Most of these involve following the course of polymerization by measuring the rate a t which the monomer, or the monomer and dimer, disappears. This in turn involves distinguishing monomer and dimer, which react rapidly with molybdic acid, from higher polymers that react more slowly. The method is particularly useful because, a s described earlier, the color-forming reaction is carried out a t a low p H where the polymerization or depolymerization of silica is a t a minimum. Hence samples taken from rapidly polymerizing or depolymerizing solutions a t higher or lower p H are frozen a t the moment they are added to the molybdic acid reagent. The monomer and dimer react very quickly, whereas each higher polymeric species depolymerizes a t a slower, characteristic rate. The structure of the silicomolybdic acid is such that within the molecule there is a tetrahedron of four oxygen atoms in which only one silicon atom can fit (20, 24). Thus only monosilicic acid, Si(OH),, can react directly. All polymeric species must first depolymerize to monomer. The silicomolybdate anion SiMo,,O,, apparently has a compact structure similar t o that established for basic aluminum chloride in

1%

Polymerization of Silica

Figure 3.8. Structure o f silicomolybdic acid. All corners of octahedra are occupied by oxygen atoms. A molybdenum atom is at the center of each octahedron. (see text).

which the polybasic aluminum ion is A I 130,(OH),,(H,0),,+, as established by Johansson (71). The analogous structure for the silicomolybdate ion requires that all the oxygen sites be filled by oxygen aioms (including those that in the aluminum complex are filled by OH groups and coordinating water molecules): SiMo,,O:;. The structure of the silicomolybdate heteropolyion is shown in Figure 3.8. The details of the sharing of the oxygen atoms (or ions) between the MOO, octahedra and the central SiO, tetrahedron have been described clearly by Cotton and Wi Ik inson (72a).
Alpha and Beta Silicic Acids

Polymers of silica were classified first by Goto (45) into two types. A reacts rapidly with molybdic acid and has a low degree of polymerization of less than four, whereas B reacts more slowly with increasing molecular weight. It appears that the difference is the size of the ultimate particles and thus the reaction rate varies in proportion to the specific surface area. Other workers have variously defined the increasing degree of polymerization, as evidenced by decreasing rate of reaction with molybdic acid, as alpha, beta, and gamma. Usually alpha is defined as silica that reacts almost completely in less than 5 min. Beta reacts completely in 10-30 min.

Characterization of Silicic Acids

I97

and has been classed as an oligomer or oligosilicic acid by Baumann (72b). G a m m a then is the higher polymers that d o not react after 10-30 min; it is often referred to simply as higher polymers. Goto and Okura (72c) proposed that the monomer and dimer species which reacted in 5 min be classed as type A . These could be removed from solution by a strong-base anion-exchange resin. They recognized that there are different types of B type polymer, since those formed in an acid medium depolymerized more rapidly than those formed in a basic medium. A method that is said to distinguish alpha and beta from gamma silicic acid was developed by Nemodruk and Bezrogova (73a), who defined the gumma silicic acid as that which did not react with molybdic acid reagent at 100C in 20 min, whereas beta reacted completely.
Measurements of Reaction Rates

A number of investigators began to use the procedure developed by Alexander (24a) to measure the rate of reaction of specific polysilicic acids with molybdic acid. This, in effect, was a measurement of the rate of depolymerization i n the colorimetric reagent. It was hoped that once the reaction rates of individual polyacids were known, the more complex reaction rate of a mixture of polymers could be interpreted as a distribution of molecular weights. The depolymerization of a particular species of silicic acid is a krst-order reaction so that the species can be characterized by a specific reaction rate constant. Since in most solutions monomer is already present along with a higher polymer or colloid, the following equations will hold:
C, C,
= =
=

M, Po
- =

silica reacted with silico-molybdate at time t total silica in the system at I = 0 total monomer in system a t t = 0 total polymer in system a t t = 0

C,

M, (1 -

+ P,(1 - e-P

I )

Ct

M, + Po

where k , and k , are the reaction velocity constants for monomer and polymer. Taking a hypothetical case where 735 ppm of monomer is in equilibrium with 7265 ppm of cubic octamer, or 9.2% of the silica is monomeric, with the known values k , = 1.5 and k , = 0.45, the color development curves are calculated from the equation. In Figure 3.9, curve A is the curve that results when all the silica is monomeric, B is the curve for the above mixture, and C is the curve for higher polymer alone. It will be noted that the amount of monomer would be difficult to estimate from this plot. However, by plotting the log of the fraction of silica not yet reacted a t time t against time, as in Figure 3. IO, lines are obtained for A and C and, a t longer times, also for the mixture B. The linear part of E extrapolates a t zero time to the fraction of higher polymer (90.8%) in the mixture.

IO

00
0

5
MINUTES

IO

Figure 3.9. Reaction o f silica S i t h molybdic acid. Calculated curves: A , monomeric silica, B . a mixture of 9 . 2 % monomer and 90.8L?c cubic octamer; C. cubic octamer alone.

10

0
W

I -

y
U

0 5

a
Ln

0 = I

a E
LL

0 2

4
E

01

0 05 0 I

3
MINUTES

Figure 3.10. Reaction of silica with molybdic acid, A , monomer, B, 9 1 % monomer and 90 0% cubic octamer. C, cubic octamrr alone

I98

Characterization of Silicic Acids

199

Because of rather low precision the method is of value only for distinguishing monomer and very low polymers from relatively high polymers, not for following the early stages of polymerization. However, Baumann (72b) studied the early stages by stopping the reaction by adding citric acid and reducing the yellow complex to the more sensitive molybdenum blue. Alexander's method (24a) was used by Thilo (73b) and several other investigators to characterize polysilicic acids by the rates of reaction with molybdic acid, each having a characteristic reaction velocity constant k. Their procedure, in slightly modified form for convenience, is given in detail in Chapter 1 as a recommended procedure. In some cases the polysilicic acid acid must be liberated from a crystalline silicate in acid at 2"C, or even i n methanol-HCI, to obtain a solution stabilized long enough to take a sample for the molybdate test. The reaction of molybdic acid with disilicic or linear trisilicic acid is rapid because these depolymerize to monomer within a few minutes at pH 3. Schwartz and Knauf (21) prepared the pure methyl esters of these acids and found that by the time they had completely hydrolyzed in water in 4 and 10 min, respectively, only monomer was present in solution. The molybdic acid was somewhat modified by Coudurier, Baudru, and Donnet (36) for their extensive study of polymerization of disilicic acid. Two solutions of molybdic acid were used containing 4 and 6 g I-' ammonium molybdate, respectively, both at pH 1.4. These contain 0.0235 and 0.0352 g-atoms I - ' molybdenum. When they were reacted with monomer at 25OC the reaction rate constants were 2.1 min-' for the more dilute and 2.6 inin-' for the more concentrated solution. However, with higher polymers the reaction rates were the same, thus indicating that the slow step is the depolymerization to monomer: polymer -Si(OH),+
ka

k,

silicomolybdic acid

However, disilicic acid also reacted at different rates with the two different concentrations of molybdic acid, indicating that it dissociates very rapidly to monomer. Equations were developed on the basis that polymer must first depolymerize before reaction. Using these equations, experimental data plotted as logarithm of unreacted silica versus time can be resolved to give the relative proportions of monomer, dimer, and polymer. The reaction rate of molybdic acid with specific polysilicate anions has been measured after obtaining a solution of the free polysilicic acid by dissolving waterinsoluble, but acid-soluble, crystalline silicates of known crystal structure. Wieker (74) applied this method to a number of calcium silicates. Four different types of silicic acid were characterized by their rates of reaction with molybdic acid, by Funk and Frydrich (75). However, they did not use the method of Alexander. Instead, the reagent was more concentrated containing 0.28 g-atoms I-' Mo with a H': M o ratio of 1.5. The high concentration of molybdic acid and the relatively low acidity accounts for the rapid reaction of this reagent with monomer and also its promotion of the depolymerization of polymers more than twice as fast as Alexander's reagent (75). I t will even gradually attack quartz. The reaction was followed not

200

Polymerization of Silica

colorimetrically, but by precipitating the silicomolybdate as quinoline salt and titrating the latter with base. Thus the method has the advantage of not requiring a spectrophotometer or colorimeter (for details see Chapter I ) . As sources of the silicic acids, crystalline acid-soluble salts of monosilicic, disilicic. and cyclic tri-, tetra-, and hexasilicic acids were dissolved rapidly in methanolic HCI, in which the silicic acids are more rapidly dissolved yet are more stable against further polymerization than in water. The liberated silicic acids were reacted at once w i t h molybdic acid reagent at 20C. For each silicic acid the reaction is first order and the constant is calculated:

where C is the fraction of unreacred silica at time 1, and K is the rate constant (C 1 .O at I = 0). Then k (min-') = 0.693 ( t h ) - l , where th = half-life.

Reaction Rate Constants


Values of constants for silicic acids from known crystalline silicates are given in Table 3. I . I t is emphasized that these apply only when Funk and Frydrych's type of reagent is used. The reaction rate decreases more rapidly than the increase in number of siloxane bonds that must be hydrolyzed to depolymerize the polysilicic acid to monomer. This is probably because of the greater stability of the ring structures as compared to corresponding chain polymers. Since several investigators have used nearly the same molybdic acid reagent solution as used by Alexander (24a), a number of values for the constants can be compared for monomer and polymers. excluding those of Funk and Frydrych, who used other reaction conditions. Each polysilicic acid i n Table 3.2 was prepared from a particular crystalline silicate known to contain that polysilicate anion, by dissolving it under conditions that avoided changing the structure. The linear polysilicic acids hydrolyze rapidly to monomer according to O'Connor (77), and the linear pentamer should have a rate constant of 0.66. that is, 90% reacted in 3.5 min. I t appears that in the case of all linear, cyclic. or polycyclic silicic acids where all siloxane bonds are exposed to the solution. the rate of depolymerizaTable 3.1. Reaction Rate Constants of Silicic Acids with Funk and Frydrych's Molybdic Acid Reagent
~ ~ ~~~

Starting Silicate

Dissolved
Cd*SIO,

Silicic Acid TY pe

Half-life (secl

? :

(Sec

K (min

'I

Si(OH),
(HO),SiOSi(OH), I(H 0 , 3 1 0 1 3

CalNa,SilO,
cd,sI,op

K,H,Si,Oi,
Cu,Si,O,, 6H20

I( HOhSlOla
(( HO),SiO],

55 23 5 360 83 0 3600

012 0032 0019 00084 00018

72 I 9 I I4 0 SO
0 II

Characterization of Silicic Acids

20 1

Table 3.2. Reaction Rate Constants of Silicic Acids with Alexander's Molybdic Acid Reagent Silicic Acid
Degree of

Rate

Polymerization
I

Type of Polymer
Monomer

Constant, K (min-I)
2.3 I .7 2.05 1.87 I .5 2. I 0.9 0.9 1.09 0.82 I .oo 0.67 0.79, 0.65 0.66 0.6 0.51

Author Alexander Thilo et al. Marsh et al.

O'Connor Hoebbel et al. Coudurier et al.


Alexander

Dimer

Cyclic Linear Cyclic Double 4-ring. cubic High mol. wt. High mol. wt.

Thilo et al. O'Connor Hoebbel et al. Cordurier et al. Thilo et al.


Hoebbel et al.

5 6
8

O'Connor Hoebbel et al.


Wieker et al.

0.46 0.42
0.050
0.015

Single chain
Double chain

Hoebbel et al. Hoebbel et ai. Wieker Hoebbel et al.

tion is so rapid that the rate of reaction with molybdate does not increase very greatly with the degree of polymerization. However, with Frydrych's faster reacting reagent, differences in depolymerization rates are more apparent.
Composition of Molybdic Acid Reagents

The compositions of the solution in which the color was actually developed are summarized in Table 3.3. One group of investigators used the Alexander composition essentially unchanged. Others modified this for specific reasons. Except for the compositions of Funk and Frydrych and of Nemodruk, the reaction rate constants with the various silicic acids are all about the same. The more concentrated reagents, such as Iler's, permit the use of higher concentrations of silica. The indicated ratio H + : Mo is not based on the H + ion concentration in the solution, but is the ratio of acid to ammonium molybdenum used in making up the mixture. Actually, a ratio of H + : M o of0.86 is required to neutralize the NH,' ion.
Other 0bservations The composition of silicomolybdic acid, determined by Khomchenko et al. (78), corresponded to H,[SiMo,,O,,]~ 29H,O. Complete conversion of silicomolybdic acid from the beta to the less intense MOO,*- of 1.66: 1.0 according to Mars yellow alpha form was observed at a H + :

202
Table 3.3. Composition of Molybdic Acid Reagents
~

Polymerization of Silica

Author
~~

H*.Mo
44 44 44 44

Mo (g-atoms I-)
0 0221 0 0227 0 0227 0 0227 0 0235 0 28 0 0600 0 0566 0 0566 0 0226

Acid HzSO, HzSO,, HCI HCI HZSO, HCI HZSO, HzSO,

Alexander
Thilo et al

Marsh et al OConnor Coudurier et al Funk and Frldrych Govett ller Kautskk Nemodruk

(PH I 4)
15
33 53

H W ,
HzSO, HSO,

65
22

(79). I t is for this reason that a H + :Mo ratio greater than 4 is generally used to develop the beta form. Sugars and other polyhydroxy organic compounds interfere with the reaction of molybdic acid with monomeric silica. This is believed to be due t o the formation of stable complexes with the molybdic acid (80). Goto and Okura (81) were the first to recognize that the depolymerization of silicic acid is catalyzed by the presence of molybdic acid. Thus at pH 1-2 in the presence of HCI alone, polysilicic acid formed monomer only very slowly, as shown by adding molybdic acid after 50 min. The rate of formation of silicomolybdate was then the same as when molybdate was added a t the start. However, it is not known whether the molybdic acid is actually involved as a catalyst by direct interaction with the polymer or whether it simply reduces the concentration of monomer in solution to such a low level that an equilibrium between polymer and monomer is displaced. A peculiar phenomenon has been noted by Iler. When a small amount of N a F is added to a polysilicic acid solution a t p H 2 it converts an equivalent amount of the silica to SiF,*-, which, when molybdic acid is then added reacts as though it were monomer. However, if the same amount of N a F is added wirh or after the addition of molybdic acid reagent it does not depolymerize an equivalent amount of silica, but instead acts as a catalyst for the depolymerization of polysilicic acids. When N a F is added before the molybdic acid so that it is converted to SiF,Z-, then when the latter reacts with molybdic acid, the fluoride ion combines irreversibly with molybdenum so that is is no longer free in the system. When added later, the molybdic acid reacts with monomer as it is developed, but does not inactivate the fluoride, which at the low p H is probably present as H F .

Separation of Silicic Acids

Although the rate of reaction of molybdic acid with individual polysilicic acid species obtained from crystalline silicates can be measured, the results are of no

Characterization of Silicic Acids

203

value in studying the polymerization reaction unless it can be shown which polyacids are actually present in the polymerizing mixture. For this reason, methods of separating the oligomers or low molecular weight species are essential. A few examples follow. Chromatography can be used, provided conditions are chosed to minimize polymerization or depolymerization during the procedure. Wieker and Hoebbel (22) found that by working rapidly, monomer, dimer, and higher species can be separated by paper chromatography in 3-4 hr using dioxane containing ( a ) 1.6 g l-lCC1,COOH and 30 g I - ' H,O to separate monomer and lower polymers, or (b) 8.0 g I-' CCI,COOH and 90 g IL,H ,O to separate higher cyclic polymers. The paper is dried and the separated spots are developed by spraying with 0.1 N N a O H and aged wet for I O min to depolymerize the silica, then with 2% ammonium molybdate in 0.3 N HCI and aged wet horizontally for 30 min, then the yellow spots are reduced to blue with 0.1 N ascorbic acid and bleached with ammonia gas to destroy molybdenum blue, thus leaving the spots of blue silicomolybdate. Low molecular weight silicic acids were separated by Baumann (82) with paper chromatography using a mixture of isopropyl alcohol, water, and acetic acid as the moving liquid and the molybdic acid reaction to locate the separate species. Polysilicic acids of different molecular weights can be separated and molecular weights estimated by gel chromatography on Sephadex columns, using 0. I M NaCl solution adjusted to p H 2 with HCI as the eluent. A blue dextran 2000 in 0.2% solution was used as a standard. Tarutani (83) made silicic acid at a concentration of 500 ppm by neutralizing the monomeric solution of sodium metasilicate with acid to p H 7. This solution was aged for various lengths of time and then acidified to p H 2 t o stop polymerization. Polysilicic acids of low molecular weight have been isolated as trimethylsilyl esters and separated by thin layer and gas chromatography by Hoebbel et al. (84). Specific polysilicate ions known to exist in certain crystals were used to make the corresponding trimethylsilyl derivatives to use as standards. This method makes it possible to separate these derivatives and characterize them further by gas chromatography and mass spectroscopy. The sources of individual silicic acids and their chromatographic constants are listed in Table 3.4. The derivatives were separated, using a mixture of Merck alumina G and Merck silica gel G as adsorbent and n-heptane as solvent. Programmed temperature chromatography was also used (84).

Particle Size and Surface Area by Titration

At a relatively early stage in the polymerization it is possible to characterize the polymeric silica, or silica particles in terms of the specific area of the silica-water interface. This is done by measuring the adsorption of hydroxyl ions in the p H range 4.00-9.00 (Beckman Type E electrode) in a nearly saturated salt solution which permits the surface charge denstiy to approach a maximum. This method was developed by Sears ( 8 5 ) t o determine the specific surface areas of colloidal particles and gels. Then it was found that if carried out rapidly it could give reproducible

Next Page
204

Polymerization of Silica

Table 3.4. Sources of Individual Silicic Acids and Chromatographic Constants of their Trimethylsilyl Derivatives

Chromatographic Constants Source Na,H,SiO,' BH,O Ca,SiO, Na,Cd,( Si,O,,) Silicic Acid Si(OH), Si(OH), OH (HO),SiOSiOSi(OH), OH
I .25
0.56

1.35

0.60

Si,O,CI,, OHOH
[( HO),SiO],

1.51

0.68

(cyclic trimer)

0.71"

0.31"

[( HO),SiO], (cyclic tetramer) Tricycloheptasilicic acid Cubic octasilicic acid IHOSiO, J8 Cubic octasilicic acid IHOSiO, ,II
a

I .oo 0.38 0.26


0.26

0.44 0.16 0.11


0.1 I

The constants for the "cyclic trimer." as compared to those of the cyclic tetramer. suggest that it is more stable and less reactive even though the trimer ring should be under greater strain Source. Hoebbel et al. (84)

results on sols of particles only 3-4 nm in diameter with a specific surface approaching 1000 mz g-I. However, in sols of such small particle size, there is an appreciable concentration of monomer at equilibrium. Also, in alkaline sols at pH 9-10.5, there is an appreciable amount of ionic silica which is converted to monomer before the titration. Since monomer reacts with base at pH 9 it is therefore necessary to correct the titration for the effect of soluble silica in order to obtain a reliable value for the specific surface area of the polymer. The term "soluble silica" is used to include the ionic silica and dimer which react with alkali-like monomer. The soluble silica can, of course, be removed a t pH 2 either by washing the silica in a filter or ultrafilter or by centrifuging. Also the sample can be adjusted to pH 8 and let stand a few hours until the soluble silica has been polymerized upon the colloidal material.

S-ar putea să vă placă și