Sei sulla pagina 1di 32

CHAPTER

Colloidal
Silica-Concentrated Sols
DEFINITION OF COLLOIDAL SILICA A N D HISTORICAL
DEVELOPMENT

The term “colloidal silica” here refers to stable dispersions or sols of discrete parti-
cles of amorphous silica. By arbitrary definition, the term excludes solutions of poly-
silicic acid in which the polymer molecules or particles are so small that they are not
stable. Such solutions, which are usually obtained by acidifying sodium silicate solu-
tions or by hydrolyzing silicon esters or halides at ordinary temperatures, have been
discussed in Chapter 3 as precursors of colloidal particles.
Stable concentrated silica sols that d o not gel or settle out for at least several
years became available i n the 1940s. after it was learned how to make uniform
colloidal particles larger than about 5 nm in diameter, stabilized with an optimum
amount of base.
When Vail ( I ) in 1925 and Treadwell and Wieland (2) in 1930 reviewed the status
of colloidal silica, only rarely could a silica sol containing more than 10% silica be
obtained; such sols were not stable toward gelling. I n 1933 the art was reviewed by
Griessbach (3), who reported that a 10% sol stabilized with ammonia was made by I .
G . Farbenindustries. A . G . I n 1941 Bird (4) patented a process for removing the
alkali from a dilute solution of sodium silicate by a hydrogen ion-exchange resin,
adding back a small amount of alkali to stabilize the silica, and concentrating by
heating to boil off water. I t is now evident that under these conditions silica particles
were grown to 5-10 nm in diameter. In 1945, White ( 5 ) patented a process of wash-
ing the salts out of silica gel made by acidifying a solution of sodium silicate,
impregnating i t with an alkaline solution, and then heating until most of the gel is
peptized to a sol. These processes generally gave sols containing 15-2070 solica, a t
least temporarily stabilized against gelling or settling out. I n 1951 Bechtold and
Snyder (6) developed the first process for making colloidal silica particles of uniform
and controlled size, and Rule ( 7 ) further defined the optimum concentrations of
alkali required for stabilization while limiting electrolyte impurities.
The history of the development and the state of the art in about 1954 was sum-
marized by ller (8). Further refinements by Alexander (9) in controlling particle size,
312
Growth and Stabilization of Discrete Particles 313

degree of aggregation, purity, and optimum concentration of stabilizing alkali led to


stable sols of particles only 8 nrn in diameter, yet containing more than 30% silica.
Stable, translucent, aqueous sols containing up to 50% by weight of SiO, have been
developed ( I O , 1 1 ) by making particles 20-25 nrn in diameter and adding an
optimum amount of alkali for stabilization and sufficient salt to reduce the viscosity
without destroying stability.
Sols containing discrete particles as large as 300 nm or more in diameter, which
settle out on standing, have been made by autoclaving wet silica gel with a base
under superatmospheric pressure and then breaking the lightly aggregated particles
apart i n a colloid mill (12a).
Thus in the past 30 years methods have been developed for making discrete silica
particles covering the whole range of colloidal size and stabilizing these as
concentrated commercial sols.
A broad review of “hydrosols” has been written by Napper and Hunter, including
their preparation and properties (12b).

GROWTH AND STABILIZATION OF DISCRETE PARTICLES

As discussed in Chapter 3 (see Figure 3.1) silicic acid polymerizes to form discrete
particles in the pH range 7-10. N o aggregation occurs if the concentration of elec-
trolyte is generally less than 0.1-0.2 N , depending on silica concentration.
Because of the nonuniform distribution of particle sizes, especially when particles
are smaller than I O nm, spontaneous particle growth occurs. The particle size that is
attained strongly depends on the temperature. As shown in Table 3.5, a t 50-IOO0C
particles reach 4-8 nm in diameter, whereas a t 350°C in an autoclave they may
grow to 150 nm. This spontaneous growth is relatively independent of the silica
concentration. The size of particles produced by autoclaving aqueous sols is limited
at high temperature by the conversion of amorphous silica t o crystalline quartz. Fyre
and McKay have shown that the rate of conversion under autogenous pressure at
330°C is proportional to the square of the hydroxyl ion concentration (13). Oehler
(14) reported that after 4 weeks a t 150°C and 2000 bars microspheres of crystalline
tridymite 50 microns in diameter were formed.

Increasing Particle Size By Adding “Active” Silica

It was because of the small particle size of the colloidal silica made a t ordinary
temperature that stable concentrated sols could not be obtained. Because higher
temperature was known t o accelerate gelling, it seemed logical that sols should be
made and kept at ordinary temperature. Hence the ultimate particle size seldom
exceeded 2-3 nm. When a sol of this type was adjusted t o pH 8-10 and vacuum
concentrated at 20-30°C it gelled when the concentration reached about 10%. It was
not realized that it could be heated and evaporated at 100°C to a stable concentra-
tion of I0-20% because a t the higher temperature the particle size increased to
4-6 nm.
314 Colloidal Silica-Concentrated Sols

However, to obtain still more concentrated sols, even larger particle sizes were
needed. This was first achieved by the further deposition of soluble silica on the
particles by adding silicic acid in the form of particles smaller than 5 nm, generally
less than 2 nm in diameter or even smaller polymer species. These are “active” in
the sense that they are more soluble and dissolve in the presence of larger particles
or “nuclei” on which silica is deposited.
It was the process developed by Bechtold and Snyder (6) that first provided stable
concentrated colloidal silica sols of any desired particle size from 10 to 130 nm in
diameter. First a 3.5% solution of silicic acid is prepared by passing a solution of
sodium silicate through a bed of hydrogen ion-exchange resin to remove sodium, and
then enough alkali is added to raise the p H to above the neutral point, using around
I % by weight of Na,O based on the silica. This part of the operation follows the
Bird process (4). However, instead of boiling down the solution directly, a portion is
heated to 100°C, which converts the silica to particles at least 4-6 nm in diameter.
Then the rest of the alkalinized polysilicic acid solution in which the particles are
less than 3-4 nm in diameter is gradually added to the hot sol as it is concentrated
by evaporation. By carrying out this addition sufficiently slowly, all the incoming
silica is deposited on the original particles, which thus are grown at a uniform rate.
Water is usually simultaneously evaporated so that the total number of particles per
unit volume remains constant, but they grow in size as the sol becomes concentrated.
In the patent literature, the initial alkalinized sol containing the silica particles which
act as nuclei has been termed a “heel,” used in the sense of a “residue” in a partly
filled container to which more liquid may be added.
In a modification of the Bechtold and Snyder process, Rule ( 7 ) started with a heel
of alkalinized sol but added a solution of polysilicic acid made by ion exchange to
which no alkali has been added. The silica particles were thus grown in a medium of
constant alkali content, so that stable concentrated sols were produced containing a
minimum o f stabilizing alkali. Albrecht ( 1 5 ) patented the optimum rate of addition
o f polysilicic acid in the Rule process for producing silica particles 45-100 nm in
diameter. By operating a similar process under superatmospheric pressure, particles
up to 150 nm in diameter have been produced (16).
Particles having a diameter of more than 100 nm have been prepared by Mindick
and Vossos ( I 7 ) by adding silica of average molecular weight below 90,000 to a sol
heel having a pH of 7-1 I and containing at least 0.1% by weight of relatively uni-
form spherical silica particles with an average diameter of a t least 30 nm. The rate
of addition of feed is maintained below a maximum according to the relation F, =
kS,C,, where F , is the maximum feed rate at any time f, in grams of silica added per
milliliter of mixture per hour, k is a predetermined rate constant equal t o about
0.005 for a constant volume process at IOO’C, S , is the specific surface area of the
particles at time 1 , in square meters per gram of silica, and C,is the concentration of
the silica in the mixture at time r, in grams per milliliter.
This method of increasing the size of silica particles is sometimes referred to as
the “buildup process.” Colloidal particles in which only the surface of the particles
consists of silica, whereas the interior is another insoluble material, can be made by
starting with suitable colloidal nuclei other than silica, using the buildup process.
Thus i t is possible to produce sols having the dispersion and surface characteristics
Growth and Stabilization of Discrete Particles 315

of colloidal silica, yet with platelike or fibrous particles, by starting with colloidal
dispersions of platelike or fibrous colloidal silicate minerals, oxides, metals, or other
materials, according to a silica-coating process developed by Iler (18). In growing
particles by this process “active” silica must not be added to the system more
rapidly than the available silica surface can take it up. Otherwise the solution
becomes so supersaturated that new small nuclei are formed and then the final sol is
not of uniform particle size. As shown in Chapter I , at 9OoC the maximum rate of
addition without nucleation is about IO grams active SiO, per 1000 m* hr-I area of
silica surface. The theoretical increase in particle size depends on the “buildup
ratio,” B,:

where W , is the amount of active silica added to the particles and Wn is the weight
of silica initially present as nuclei. Obviously, the particle diameter will be increased
from the initial size dl to the final size d, in accordance with the following equation:

A process by which 15% sol can be made directly by ion exchange was developed
by Iler and Wolter (19), whereby a heel of water or dilute sodium silicate is heated
and stirred and then wet, drained, regenerated hydrogen ion-exchange resin
(preferably of the weak acid type) and relatively concentrated sodium silicate solu-
tion are added simultaneously at such a rate as to maintain the pH at around 9.
Depending on temperature and relative amount and size of particles in the heel, the
rates of addition of resin and silicate can be regulated to increase the size of the
silica particles without further nucleation. A similar process is operated in a column
or fluidized bed by which resin is continuously added at the top and withdrawn at
the bottom in counterflow to the silica sol (20).
In a study of the growth of silica particles by adding monomer to a dispersion of
preformed nuclei, Bartholin and Guyot (21) studied the formation of nuclei when
low molecular weight silicic acid, made by ion exchange from sodium silicate, was
aged at pH 8. The very pure sodium silicate was made from ethyl orthosilicate. In
this case the 2.5% sol alkalinized to pH was stored at 4°C to minimize spontaneous
particle growth and, because of the high concentration of small particles, promote
formation of microgel. When heated, the microgel aggregates condensed and shrank
to more or less dense spherical particles around 100 8, diameter. However, it should
be noted that when the heating step from 4 to80”C is slow (2 h r ) the nuclei are much
more uniform in size than when the heating is rapid (IO min), as shown by the
greater uniformity of the final sol obtained by adding freshly made silicic acid solu-
tion also adjusted to pH 8.
316 Colloidal Silica-Concentrated Sols

While particles smaller than about 20 nm are being grown by this process, there is
some simultaneous spontaneous growth. This is important in making particles
smaller than 15 nm since even i f no active silica is added, some growth of nuclei
would occur over the period of several hours normally required in the buildup
process.
The above remarks apply to systems where the silica is added a s “active” silica
prepared as a separate solution, usually by ion exchange. More rapid growth occurs
when the silica is added directly a s sodium silicate to a “heel” of nuclei from which
sodium ions are being constantly removed by hydrogen ion-exchange resin, as dis-
cussed later. In this case, the silica is initially present as monomer and oligomers
and thus no time is required to depolymerize higher polysilicic acids which are
invariably present when the silica is prepared first as a separate solution of “active”
silica.
Further variations in the “buildup” process are noted as follows. Albrecht further
defined the maximum concentration to which sols of different particle sizes may be
concentrated while the size is being increased by the deposition process (22) to 100
nm. By operating the growth process above IOOOC a t steam pressures of 5-100 psi,
sols art: obtained that can be concentrated to 5 5 % 50, according to Klosak (23).
The process should make particles of larger size than obtainable at 100°C. Bartholin
and Gubot (21) studied particle growth in alkalinized solutions of pure silicic acid,
following the size of the growing spherical silica particles by light scattering. For an
increase i n concentration when no further particle growth is desired, Reven and
Blake (24) propose boiling down a sol of 28-38% SiO, particles of 13-50 nm
diameter while adding a sol of the same composition until the concentration reaches
5 2 % sic),.
lrani made particles of very uniform size by the “buildup” process by starting
with a “heel” of water or a sol of uniform small particle size a t 60-150°C and
adding silicate and hydrogen ion-exchange resin under prescribed conditions (25).
Another process variation (26) involves using a sol of “active” silica as it comes
from the ion exchange at p H 2 . 5 - 5 . 5 instead of making i t alkaline before adding it
to the ”heel sol” in the evaporation. However, it appears that the SiO,:Na,O ratio
would continually increase during the process. Sippel (27) claims that by operating
in the above manner, feeding silicic acid to a “heel sol” of 8-20 nm in diameter, and
maintaining the p H a t 8.5-9.5 by adding a solution o f sodium silicate, which thus
maintains the SiO,:Na,O ratio approximately constant, a sol of low turbidity is
produced. Starting with a “heel” of water and adding active silica is another way of
operating proposed by Weldes, Boyle, and Bobb (28). Specific conditions for making
nuclei and adding silicic acid to obtain particles over 50 nm in diameter are disclosed
by Cummings (29). Egbregt (30) proposed circulating a dilute sodium silicate solu-
tion through a column of hydrogen ion-exchange resin and adding more sodium sili-
cate at a specified rate.
As pointed out in Chapter 1, silica can be deposited on many different substrates.
Once the surface of a particle is covered with a monolayer of SiO,, subsequent
deposition is silica on silica in all cases. For example, particles of lead chromate can
be coated with an impervious layer of silica (31). Also, as described previously, coat-
ing TiO, pigment with SiO, is conducted on a large scale.
Growth and Stabilization of Discrete Particles 317

Methods of Making Particles Under 10 nm In Size

Initial commercial sols were made with particles larger than about 8 nm so they
could be concentrated to at least 30% SiO, for economical shipment. Such sols are
extremely stable in regard to further growth of particles at ordinary temperature (see
Figure 3.52). Then it was realized that for some uses smaller particles were to be
preferred and such sols at lower silica concentrations were developed. However, the
problem arises that particles smaller than about 8 nm and certainly those smaller
than 5 nm grow spontaneously in storage with corresponding changes in properties.
Furthermore, since such sols are generally concentrated to at least 1570, those
smaller than 5 nm also may undergo a certain amount of aggregation in storage with
definite changes in use properties. A number of parents relate to the production and
stabilization of sols of this type.
The earliest salt-free sols of 5-8 nm particles were made by Alexander, who
defined the silica concentration limits and alkali content for stabilization in terms of
particle size (9). The particle size was related to average specific surface area of
silica within the following limits:
Particles

Specific Diameter SiO, Ratio


Surface (m? g-l) (nm) (%) SiO,:Na,O

600 4.6 15-20 20-200


3SO 7.9 15-34 34-340

A method of making a sol of 10% SiO, with particles about 3 nm in size, devised
by Iler (32). involved starting with a sol less than 3 nm in size, adding sodium sili-
cate in an amount not exceeding 4 % SiO, of added silica, so as not to exceed a
sodium ion concentration of 0.4 N , aging 10 min, and passing the sol through a
column of strong-acid ion-exchange resin i n hydrogen form to recover a sol of pH
3.5, and repeating the process at ordinary temperature.
To increase silica concentration without particle growth beyond 5-10 nm size. a
silicic acid solution is alkalinized with sodium silicate and then more silicic acid is
added while water is removed by vacuum distillation. Under these circumstances the
low temperature avoids excessive particle growth in spite of the addition of active
silica to the system (33). An earlier approach was to form the particles to the desired
size in dilute solution and then concentrate by evaporating water and adding more
dilute sol. Since the added particles are the same size as those in the evaporator, no
“buildup” occurs (34).
The use of ammonia as a stabilizer apparently minimizes particle growth. Weldes
and Derolf (35) claim that sols of particles size 2-3 nm are obtained in concentra-
tions up to 12% when stabilized with both NaOH and NH,OH, the (NH,),O:Na,O
ratio being 25-150 and the SiO,:(NH,),O 2-8. Possibly the adsorption of NH, on
the silica surface reduces the rate of dissolution of silica so that growth is retarded.
Birkheimer (36) similarly claims stabilizing a sol with excess ammonia and then
boiling out the excess.
318 Colloidal Silica-Concentrated Sols

Particles estimated to be 1-3 nm in size were made by McNally and Rosenberg


(37) b) deionizing sodium silicate solution with a weak-acid hydrogen ion-exchange
resin which had first been neutralized in suspension at pH 6-7. This presumably
prevented the formation of larger nuclei o f silica that might be formed at lower pH
in contact with the resin. The SiO,:Na,O ratio of the sols was 4 : 1 to 40: I and the
concentration 15-25% SiO,. I t is possible that in the absence of any larger nuclei,
the sol was unusually stable against spontaneous particle growth, although a
concentration as high as 25% suggests that in that case the particles must have been
larger than 3 nm.
Vossos claimed sols o f particles less than 5 nm in size stabilized a t p H 9-1 1
containing up to 25% SiO,. The alkalinized dilute sol was vacuum concentrated a t
less than 150°F to minimize growth (38).
Reuter and Reven prepared particles 7.5 nm in diameter from a sol of 1-3 nm
particles by heating a t pH 8.7 and carrying out a “buildup” at 165’F to get a 15%
sol (39).
Marotta (40) made 5-10 nm particles by adding silicic acid solution to a dilute
solution of sodium silicate over a period of hours at 25-5OoC, then more at
5O-9O0C, and still more at 70-I00”C while evaporating water. This apparently
involved a combination o f additional nucleation and particle buildup.
Other preparations of sols with particles under 5 nm in diameter and their stabi-
lization have been discussed under the heading o f polysilicates and silicates. Sols of
such small particles stabilized with relatively large amounts of base assume some of
the characteristics of soluble silicates or polysilicates rather than of colloidal silicas.

Stabilization Against Particle Growth (Theory of Paul C. Yates)

As indicated i n Chapter 3, polysilicates or sols of very small particles are stabilized


by having sufficient alkali in the system. Yates (41) proposed that alkali-stabilized
sols are stable not only against gelling, but also i n the thermodynamic sense. He
pointed out that there are thermodynamic factors that prevent the spontaneous
growth of particles or their aggregation and stabilize the high interfacial solid-liquid
interface in the silica-water system. The main factor that counterbalances the free
energy change involved in loss of surface area in the silica-water system is the strong
adsorption of the continuous liquid phase or of stabilizing counterions or of other
adsorbed species at the surface of the dispersed phase, thus changing the free energy
of the interface.
The effect can most easily be understood by considering the changes in free energy
that occur during the following hypothetical, reversible, isothermal steps in the
three-component system Si0,-H,O-NaOH starting with a n alkali-stabilized sol:

I . Desorb N a O H from the surface of the colloidal particles: AFl.


2. Desorb water from the surface of the colloid; that is. the free energy of wetting
the surface: 1 F 2 .
3. Decrease the surface area of the colloid to the minimum possible geometrical
area allowed by its density; that is, the free energy of the solid colloid: A F 3 .
Growth and Stabilization of Discrete Particles 319

4. Return solvent to the surface of the collapsed colloid of minimum area: AF,,
5 . Return the N a O H to the system: A F 5 .

By definition,

AF, + AF, = free energy of adsorbed N a O H on colloid

AF2 + AF3 + AF, = interfacial free energy of colloid in a


binary SO,-H,O system
If
AFi + AFS = AFZ + AF3 + AF, = 0
then the initial colloid system of three components is at thermodynamic equilibrium.
In other words, if the free energy of adsorption of sodium hydroxide on the surface
of the colloid equals the free surface energy of the system in the absence of N a O H ,
then the sol is thermodynamically stable.
Heretofore all colloid systems such as sols have been considered to be thermody-
namically unstable.
In a silica sol the interfacial free energy of amorphous silica versus water is of the
order of 50 ergs cm-* (Figure 3.32). The the decrease in free surface energy
accompanying the loss of surface, assuming about 8 silicon atoms nm-*, amounts to
900 cal g-mole-' of surface silicon atoms. I f the colloidal particles are stabilized
thermodynamically, the counterbalancing free energy term must be of about this
magnitude. This stabilizing energy comes from the adsorption of OH- and the
counter N a + ions on the colloid surface.
Yates calculated the amount of alkali necessary to stabilize silica particles against
growth a s follows.
The free energy change connected with the ionization of surface groups depends
on the acid dissociation constant which, in turn, varies with the salt concentration
and the degree to which the ionization of surface groups has already occurred. The
free energy change under constant conditions will be

AF = a ( - R T l n K') (1)

where a is the fraction of surface groups ionized and K ' is the instantaneous value of
the dissociation constant at that particular value of a, other conditions being fixed,
and assuming a surface containing one mole of silanol groups.
As will be discussed in connection with the charge on particles, Yates derived a
formula for the value of K' as a function of pH, sodium ion concentration in solu-
tion, and a, the degree of dissociation of the surface.

1-a
pH = pK - n log -
CY
- 0.74 log(A N a + )

where K is the overall dissociation constant and ( A N a - ) is the activity of the


sodium ion times the sodium ion normality.
320 Colloidal Silica-Concentrated Sols

From titration data at a fixed concentration of N a * ion, a plot of -log [(I -

( i ) / n ]against pH gives a line of slope n and intercept pK. The value of n from the
experimental data is 3.47.
The corresponding expression involving K ' is

where

I - n
pK' = pK - ( n - l)log - - 0.74 log(A Na') (4)
a

From equations 1 and 4,

I-cu
-log K - ( n - l)log --(Y
0.74 log(A

I t can he shoun that,

where r is the weight ratio of 50,:Na,O in the system and S is the specific surface
area of the silica in square meters per gram: Yates assumed the density of silica to
be 2 . 3 g C I T - ~ .
Substituting the numerical values for K , a and n gives

rS - 2430
AF = 3'38 Io' [12.08 - 2.47 log - 0.74 log(A Na+)] (7)
rS 2430

which expresses the free energy change associated with moles of surface groups at
any level of ionization or salt content.
I n order for a silica sol to lower its specific surface free energy by particle growth
or aggregation to form a sol of lower surface area, it is necessary t o reverse the ioni-
zation and return the adsorbed ions t o the intermicellar liquid. According t o Yates
the free energy change will be yS, where y is the value of the specific interfacial free
energy between the silica surface and water. The latter is a function of particle size
and thus of S.
Let the free energy change occurring when particles of diameter D lose A S cm2 of
surface be
Growth and Stabilization of Discrete Particles 32I

A similar expression for a flat surface is

This is the difference in free energy between a colloid of particle diameter D and the
same colloid i f it had the free energy of a flat surface. This would be - R T
In(SD/S,), where S, and So are the corresponding solubilities of the curved and flat
surfaces of silica.
Yates used the equation published by Iler in 1955 (8):

where D is in millimicrons. This was based on an interfacial energy of 80 ergs cm -?,


However, as shown in Chapter 1 (see also Figures 1 . 1 0 ~and 3.32), the value is more
likely to be about 50 ergs cm-l so that the equation should probably be

Then

2.4
AFD - AFO = -RT -D

From equations 8 and IO,

RT
70 = yo + 3.8 -
DAS
RT
AFD -7; A S + 2.4 -
D

However, D can be related to S by taking into account the density of silica, assumed
by Yates to be 2.3 g c m - ?

2554
D = -
S
322 Colloidal Silica-Concentrated Sols

Then
2.4 R T
-AF,, = yoAS + -2554
S

Since at 25°C. T = 298°K. R = 1.987 cal deg-l mole-' and

Assuming yo = 50 mg ergs-z, AS equals the area of 6 x 10'' surface silicon atoms.


Assuming 8 silicon atoms per square nanometer, the value of S is 7.5 x IO' cm2.
Changing ergs to calories gives

-AFD = 50(2.39 x l O - ' ) ( 7 . 5 x IO') + 0.56s (19)


- AFo = 896 + 0.56s cal
Then - A F D from equation 19 must equal A F from equation 7:

3.38 x 10' rS - 2430


896 + 0.56s = 12.08 - 2.47 log - 0.74 log(A Na') (20)
rS 2430

Rearranging gives

3 . 3 8 x IO6 (12.08 - 2.47 log [(rS - 2430)/2430] - 0.74 log(A Na+)l


r = (21)
S(896 + 0 . 5 6 s )

I n the absence of foreign salts. the concentration of N a + can be calculated simply


from the concentration of silica. C,, and the ratio r since

Since the sodium concentration in stable sols seldom exceeds 0.1 N , we may assume
the activity is about unity:

where C is the molar concentration of SiO, in the system:

3 . 3 8 x 10' (12.08 - 2.47 log [(rS - 2430)/2430] - 0.74 log(ZC/r)}


r = (22)
S(896 + 0.56s)
Growth and Stabilization of Discrete Particles 323

For specific examples, assuming values for S and C. the following values of r were
calculated from the equation.
For a 15% sol of 4.5 nm particles (600 m2 g-I) the calculated SiO,:Na,O ratio is
47: I . For a 39% sol of 13.4 nm particles (200 m2 g-I) the calculated ratio i s 170: I .
(Yates’s earlier calculations gave 32 : 1 and I I5 : I .)
In commercial sols corresponding ratios of about 25: 1 and 100: I are actually
used for these particle sizes. As shown in Chapter 2, particles of about 16 nm
diameter stabilized a t 100: I ratio grow very little in 20 years. However, particles
4-5 nm in diameter in some cases undergo a slow I0-20% increase in diameter in a
year or so even though stabilized at 25: I ratio. However, lot-to-lot variations can
occur, probably because of variations in the initial particle size distribution.
The above formulas probably d o not apply to particles much under 7-8 nm in
diameter since the increasing solubility of the particles involves the presence of
appreciable concentrations of silicate ions which must be taken into account in the
stabilized systems. This would probably require more stabilizing alkali than the
present equations indicate, that is, lower values of r .

Stabilization Against Aggregation

Concentrated silica sols are stabilized against interparticle siloxane bonding by


either ( I ) an ionic charge on particles so that particles are kept apart by charge
repulsion; or (2) an adsorbed, generally monomolecular, layer of inert material
which separates the silica surfaces t o an extent that prevents direct contact of silanol
groups. This has been referred to as “steric” stabilization.
In the case of particles of appreciable size, and especially a t low pH where
siloxane bond formation is slow, spontaneous interparticle bonding is usually not
observed. Thus with particles more than 100 nm in diameter such bonding does not
appear to occur even in concentrated sols over the whole p H range unless the sol is
dried. With such large particles, even if some siloxane bonds are formed a t the
points of contact, they are probably insufficient to withstand the mechanical strain
involved when a pair of such large particles collides with a third particle in the
course of Brownian motion.
Another form of instability in storage is sedimentation by gravity whereby a very
concentrated viscous layer is formed a t the bottom of sols of large particles. In some
cases there appears t o be a distinct liquid-liquid boundary suggesting that a
concentrated coacervate has been formed. Observation of a series of 30% sols stabi-
lized a t SiO,:Na,O ratios around 100-200 showed that over a 20-year period sedi-
mentation was pronounced only when the particle diameter was over about 70 nm.
Sols of smaller particles probably remained essentially homogeneous because of
convection currents since the storage temperature fluctuated between 20 and 30°C.
The role of convection in the formation of stratified layers in silica suspensions has
been studied by Baruch-Weill (42).
There is one type of instability that is not encountered in this system, namely,
crystallization. As pointed out by Walton (43), the higher the degree of supersatura-
tion the smaller the size of the critical nucleus, which may be so small as not to cor-
324 Colloidal Silica-Concentrated Sols

respond to a particular crystalline phase. This solid phase then grows with no
particular order t o the structure. For simple compounds the ions or molecules can
rapidly rearrange to satisfy minimum energy requirements. In the case of silica the
energy difference between the amorphous and crystalline states is small and, further-
more, a high energy of activation is required to break siloxane bonds t o permit the
rearrangement. N o crystalline form of silica has ever been found in sols or gels even
after aging for 25 years at ordinary temperature.

Stabilization B y Ionic Charge

An ionic charge on the particles in the presence of alkali is the chief mechanism of
stabilization in commercial sols. However, a completely satisfactory theory of stabi-
lization has apparently not yet been developed. The basic principles of stabilization
by the ionic double layer around particles were developed by Derjaguin and Landau
(44) and Verwey and Overbeek (45), hence the ”DLVO” theory; it has been specifi-
cally applied to spherical particles (46a). An excellent summary of the forces affect-
ing the stability of disperse systems was presented by Ottewill (46b).
Napper has written a summary of colloid stability (47), including the principles, of
both electrostatic and steric stabilization. A fundamental study of the van der Waals
forces between amorphous SiO, surfaces was carried out by Roweler (48), who
measured the attraction between two fused silica surfaces covered with thin films of
chromium metal in a high vacuum. However. it is difficult t o translate these results
to an aqueous silica system. The Hamaker constant in many colloid systems has
been reviewed by Visser (49) including the SiOz-HzO system.
At least under some conditions the stability of silica sols is in complete contradic-
tion to the “DLVO” theory, as pointed out by Kitchener (SO). Thus a t p H 2, where
the charge on silica particles is zero, the particles aggregate least rapidly and the sol
has highest temporary stability. However, it is only in alkaline solution where the
particles are highly charged that sols are permanently stable. Here the double layer
theory is more logically applicable.
Matijevic ( 5 1 ) has discussed the applicability of the “DVLO” theory to various
inorganic sols. In the case of silica sols, the nature o f the electrolyte is of major
importance. Cations vary so enormously in their adsorption and formation of stable
complexes with the silica surface that the theory is of little practical value. The same
conclusion was reached by Webb, Bhatnagar, and Williams in regard to colloidal
TiO, ( 5 2 ) .

MAXIMLM CONCENTRATION OF SOLS. The problem of stabilizing sols becomes


acute at high silica concentrations, which a r e needed for economic shipment. No
appreciable change in properties should occur in less than a year under ordinary
storage conditions.
Commercial sols are presumably stabilized a t optimum p H and are concentrated
to the maximum concentration permitted by the particular particle size. I t is likely
that in all cases the particles are concentrated until the atmospheres of stabilizing
counterions surrounding the particles begin t o overlap so that a slightly higher
Growth and Stabilization of Discrete Particles 325

concentration would lead to a marked rise in viscosity and long-term instability, that
is, gelling.
The concentrations and particle sizes of typical commercial sols are plotted as
curve A in Figure 4. I . Based on the density of amorphous SiO, of 2.2 g ml-l, curves
B and C can be calculated, expressing the silica concentration as grams SiO, per 100
ml of sol and as volume of solid SiO, i n milliliters per 100 ml of sol, respectively.
The latter volume percent divided by 100 gives the “volume fraction,” c$~,of solid
silica in suspension.
Since all the sols are approximately at maximum practical concentration, the
particles plus their surrounding ion atmospheres are presumably at about equal
packing density. This maximum packing density is proportional to the volume
fraction of the dispersed phase d,, where the dispersed phase is defined as the
particle plus its ion atmosphere and d is particle diameter

where K is a constant, and d, is the thickness of the ion atmosphere or Gouy double
layer around the particle. Rearranging gives

Plotting 4,’I3/d versus 4s1/3from the data of curve C of Figure 4.1 results in the
linear plot of Figure 4.2. From this, d , = 2.43 nm and K = I .68.

For large particles, where the thickness of the ionic layer is much smaller than d ,
4s= l / K = 0.6. This indicates the particles in such sols are packed somewhat more
closely than random packing.

- Figure 4.1. Maximum concentration


versus particle size in stable aqueous silica
sols at about pH 9.5. A , concentration
(wt. a); B, concentration (grams SiO, per
0 IO 20 30 40 50 60 100 ml); C, volume fraction of S O ,
PARTICLE DIAMETER, (nrn) (X 100).
326 Colloidal Silica-Concentrated Sols

The thickness, d,. of the Gouy double layer in a given system depends only on i,
the ionic strength (53).

d, = 3.051'~ A

Since it is calculated for the commercial sols that d,, averages about 24 A, the
mean ionic strength calculated from the equation would be about 0.016 N for all the
sols. Actually. in a typical 15 nm sol containing about 500 gl-' of SiO,, the usual
stabilizing ratio of SiO,:Na,O is about 100: I and the Na,O content is 5 g I - ' or
about 0.16 N . In such concentrated sols most of the sodium ions must lie very close
to the surface ( i n the Stern layer) leaving a concentration of the order of 0.016 N in
the outer double layer.
The amount of alkali required for stabilization is generally proportional to the
surfacc area o f the silica particles in the sol. I t amounts to roughly 1.0-1.5 molecules
N a O H nm-, in most sols which may contain IO-, N salt impurities. However, i f the
sol is nearly free from salt (less than N ) then the concentration of stabilizing
N a O H can approach the theoretical value of 0.016 N giving a surface concentration
of 0.3 molecules N a O H nm-,.

Addition o/Salr to Lower Viscosity

In sols with particles larger than IO nrn in diameter which can be concentrated to
500-700 g I - ' S O , , the viscosity becomes very high. I t can be lowered by a minor
reduction in the thickness of the double layer by adding a small amount of salt
without making the sol unstable. Atkins was the first to use this technique (54). He
disclosed stabilizing salt-free sols of particles in the range 15-40 nm in diameter with
alkali and adding a salt such as Na,SO, or N a C l to give a concentration of
0.01-0.04 N . This made it possible to concentrate the sols to considerably higher

0 Figure 4.2. Relationship between volume


0 0 05 0 10 fraction ($J$) of silica and particle diameter. d
(plotted as 3/d), i n stable silica sols ot
c$s"5d
maximum concentration.
Growth and Stabilization of Discrete Particles 327

silica concentrations without the excessive increase in viscosity which occurred when
no salt was added. Later VOSSOSand Mindick (55) added a t least 0.003% by weight,
based on silica, of a metal-free ammonium or organic base salt of an anion of a very
weak acid. Thus NH,HCO, is added t o reduce the viscosity of a concentrated sol
without introducing any impurities that would be nonvolatile when the silica was
dried. With the same objective of reducing the viscosity of a concentrated sol,
Marotta (56) claimed that a combination of 0.055-0.095 N Na,SO, plus enough
sodium silicate to give a p H of 8.8-9.9 gave a stable sol but if N a O H is used to
adjust pH, gelling occurs.

Steric Stabilization

The theory of steric stabilization of sols has been discussed by Smitham, Evans, and
Napper (57) and Bagchi (58). There have been some partly aqueous silica sols
developed that are stabilized solely by adsorption of nonionic molecules. A typical
example is the 35% sol of Luvisi (59) which is salt-free and adjusted to pH 3-4.5,
containing 30-9070 of a monohydric alcohol, for example, isopropanol. At pH 4-4.5,
glyoxal apparently has a stabilizing effect on silicic acid sols and at the same time
promotes the insolubilization of water-soluble polymers such as gelatin or polyvinyl
alcohol (60). I t is well known that silicic acid sols free from salt at low pH are more
stable in the presence of hydrogen-bonding agents such as lower polyether-alcohols;
no doubt, more concentrated sols of colloidal particles could be similarly stabilized.
However, maximum stability is attained when all water has been removed. Such
sols, generally referred to a s organosols, are covered later in this chapter. A typical
process involves complete removal of cations and anions from a silica sol leaving the
p H between 1 and 7 and then partial or complete replacement of water with an
alcohol, ether, or ketone (61).
Steric stabilization has been used in the case of very small particles to supplement
ionic stabilization. Thus Yates stabilized sols of very small particles with a combina-
tion of an inorganic or organic base with a water-soluble nonaromatic polyhydroxy
or hydroxy-ether compound, for example, polyvinyl alcohol (62). There is probably
also some steric stabilization when an organic base cation such as (CH,),N+ is
present since, as Wolter (63) found, a silica sol of this type can be evaporated to a
dry powder that will spontaneously redisperse in water. Such sols also can be redis-
persed after freezing.
Steric stabilization by a monolayer of hydrogen-bonded water molecules (64)
must be involved in the case of sols which appear stable a t least for several months
at low pH. Extremely low concentration of ionic impurities, other than the necessary
acid to maintain the low pH, is essential. Alkali tends to be buried within particles
and comes out slowly, raising the p H and causing gelling unless special precautions
are taken. Mindick and Reven (65) repeatedly deionize, age, and deionize the sol
with mixed cation- and anion-exchange resins until the pH does not rise above 3 dur-
ing storage. It must be pointed out that traces of aluminum are present in all com-
mercial silicates and may still be present as S i A I 0 , - H + sites on the silica surface in
sufficient amounts to contribute to the stability a t p H 3.
328 Colloidal Silica-Concentrated Sols

Porous Particles

The primary particles of colloidal silica are generally nonporous if formed or grown
in alkaline solution and especially if formed at elevated temperature, either above
6OoC in aqueous solution or condensed from the gas phase a t very high temperature.
Fused silica (glass) has a density of about 2.20 g c w 3 . I f one measures the density
of colloidally subdivided amorphous silica powders by immersion in xylene and cor-
rects for the small content of surface O H groups. assuming the equivalent H,O to
have a density of 1.0. the following densities have been observed:

Silica Density (g c m - 3 )

Commercial sols 2.2-2.3


Commercial gel (large pore) 2.22
Condensed from vapor phase 2.16
Precipitated from hot solution 2.0

In all these powders the surface area determined by nitrogen adsorption is in


approximate agreement with the area calculated from the distribution of particle
sizes observed in electron micrographs, assuming a density of 2.2 g c m - 3 (66). From
this it can be concluded that particles of this type have a porosity to nitrogen of less
than 5-10 vol. %.
On the other hand. silica sols can be made i n which the discrete particles are quite
porous. They are formed generally a t ordinary temperature in water by the uniform
aggregation of much smaller ultimate particles, less than 5 nm in size. Radczewski
and Richter (67) obtained porous 200 nm particles by hydrolyzing SiCI, and aging
the neutralized solution. Stober, Fink, and Bohn made similar particles by hydrolyz-
ing ethyl silicate in a mixture of ammonia, water, and alcohol (68). The most
detailed study has been made of porous silica particles prepared by the dissolution of
powdered silicon metal in water. Balthis (69) discovered that tine silicon metal
powder, washed with dilute H F to remove the oxide film, will react with water with
evolution of hydrogen in the presence of an organic base catalyst to give clear dilute
sols of particles 8-15 nm in diameter that can then be concentrated by evaporation.
These particles were nonporous. Balthis then found that in the presence of ammonia
rather than an organic base, large particles were formed, u p to 35 nm diameter.
Furthermore, the particles were shown to be porous. They depolymerized or
dissolved in dilute alkali at a more rapid rate than particles of the same diameter
that were known to be nonporous. The total surface area, measured by adsorption of
nitrogen (BET method) was t w o or three times the external area measured by
adsorption of methyl red dye from benzene solution.
McMillan then further studied the reaction and made particles larger than 100 nm
in diameter which were highly porous (70). By changing conditions during the reac-
tion, the porosity of the inner and outer regions of the particles could be changed
and the size could be made t o range from I00 to 500 nm. Spherical particles of very
uniform diameter could be obtained. A study of the particle structure, by measuring
Growth and Stabilization of Discrete Particles 329

the rate of depolymerization (dissolution) in alkali, indicated the constituent primary


particles were 4 nm in diameter.
The freshly made particles obtained by reaction at 25-30°C were highly porous to
nitrogen. the density being about 1.3 corresponding to a porosity of 40% by volume,
Spherical particles 200 nm in diameter with an external specific surface area of 15
rnz g - l resulted after the sol had been dried to a powder. However, before drying, the
internal surface area, measured by the adsorption of O H - ions, was 745 m2 g - l , cor-
responding to ultimate or primary particles about 4 nm in diameter.
When a sol of such porous particles was heated at 90°C, pH 9.8, the area
determined by nitrogen adsorption decreased to 22 m2 g-l, indicating that the
diameter of the internal pores had decreased so that nitrogen molecules could not
penetrate and thus could be adsorbed only on the outer surface. However, micropo-
rosity still existed because the area determined by O H - ion adsorption was still 707
mz g-', By heating the sol for a longer time a t pH IO, the pores could be further
closed and at least the surface became completely impervious. Some water and
alkali were no doubt trapped within.
Colloidal particles which are probably porous were made by Cummings (29) as a
milky sol when a solution of silicic acid was fed into a hot solution of sodium silicate
containing 0.35-0.4 N sodium ion. Until the sodium ion concentration is reduced by
dilution by the added silicic acid, the latter must be partially aggregated by the
sodium ions so that deposition of silica on small aggregates produces a sol of parti-
cles ranging from 20 t o 250 nm in diameter.
Production of a sol of porous colloidal particles was observed by Her when a solu-
tion of sodium silicate, S i 0 2 : N a 2 0 = 3 . 2 5 : 1.0, diluted to 4% SiO,, was heated to
9OoC and then 60% of the alkali was slowly neutralized by adding H2S0,.The high
temperature was necessary because when neutralization was conducted a t 60°C only
a soft translucent gel was obtained. The mechanism of the process appears to be as
follows.
In the initial 3.25 ratio sodium silicate solution about a third of the silica was
present as colloidal particles about 1-2 nm in diameter and the remainder as
HSiO; ions, When part of the alkali was neutralized, the silica from the H S i 0 , -
ions was deposited on the colloidal particles which at the high temperature increased
in size to 3-4 nm. At the same time, the sodium sulfate concentration was 0.3-0.4
N , which was above the critical coagulation concentration SO that the highly charged
particles nevertheless aggregated into porous spheres 15-20 nm in diameter. The fact
that the spheres were made up of 4-5 nm dense silica particles was indicated by a
surface area of 600 m z g - l as measured by adsorption of O H - ions before the
spheres were dried. After being dried the surface area by nitrogen adsorption was
only 160 m2 g-' corresponding to dense 17 nm spheres. T h e initial spherical
aggregates therefore shrank and the constituent ultimate particles became tightly
packed together, eliminating the internal porosity.
It is probable that if the initially formed wet particles had been acidified t o pH 2
and washed with propanol before being dried, much of the internal porosity would
have been retained.
Several stages of aggregation of smaller silica particles into larger ones and these,
330 Colloidal Silica-ConcentratedSols

in turn, into even larger ones have been described by Greer (71). Stages of particle
sizes I .5. IO, 60, and 331 nm were described and an equation was given:
d = Or"-'

where d is the diameter of the sphere at stage growth n , a is the diameter of the units of
the initial growth stage, and r has a value of6. I f one assumes that a is 1.5 nm, then the
stages should occur at 9, 54. and 324 nm. Conditions for bringing about such aggrega-
tion were not disclosed.
Microporosity in colloidal particles may in some cases be demonstrated by low
angle X-ray diffraction. When the particle size determined in this way is much
smaller than the size calculated from the specific surface area by nitrogen adsorption
or observed in electron micrographs it means that the particles are made up of still
smaller discrete units so closely packed that the pores are not large enough'lo be
penetrated by nitrogen molecules (72). Most silica gels consist of primary particles
with intervening pores accessible to nitrogen. However, Lederer, Schurz, and Janzon
(73) reported that particles in their particular silica gels appeared to have some
"inner" area, because hydration corresponded t o 0.15-0.26 gram H,O per gram of
silica, which would indicate a high porosity.

Elongated Particles

An astounding observation of elongated or spindle-shaped silica particles has been


reported by Ogino and Kuronuma (74). This is the first time that discrete silica
particles of colloidal size have been observed with a shape other than spherical. They
were formed by adding hydrogen iqn-exchange resin slowly to a dilute solution of
sodium silicate (3.52 SiO,) at 40°C until the pH dropped to 3. then adding ammonia
to raise the pH to 8-9 and heating 1 h r at 80°C.

Particles With Non-Siliceous Cores

As long as the surface of a colloidal particle consists of silica, its colloid properties
are the same as those of solid silica particles. As discussed in chapter 1, silica can be
deposited on a wide variety of surfaces, as shown by Iler (18). T o coat sol particles
carrying a positive charge, like Fe,O, or AI,O,, it is first necessary to reverse the
charge by adding the dilute sol into a dilute (10%) solution of sodium silicate under
conditions of intense agitation. I t is also possible to reverse the charge with a chelat-
ing agent such as citrate before adding silicate. The surface is thus covered with a
negatively charged molecular layer of adsorbed silicate on which a layer of SiOpcan
be applied. A sol of silica-clad thoria particles has been made by Barrett, Moises,
and Vanik (75). Stable sols of particles greater than 50 nm in diameter contained up
to 60% solids. The silica coating was about 50% by weight of the thoria core. Silica-
coated particles of thoria and urania are claimed by Fitch, Sanchez, and Vanil (76).
The kinetics of absorption of silica onto the surface of particles of thoria was
determined at 100-200°C in an autoclave (77).
Methods of Making Sols 33 1

METHODS OF MAKING SOLS

Many processes have been proposed and are employed for producing colloidal silica
from low-cost sodium silicate solutions. Sols have also been made from hydrolyzable
compounds such as ethyl silicate or silicon tetrachloride. To produce sols that are
stable at a reasonably high concentration, it is necessary to grow the particles to a
certain size under alkaline conditions where the particles remain negatively charged
so they will not flocculate or gel. The particles should also be nonporous.
Colloidal silica is formed in nature when water i s saturated with quartz at high
temperature and pressure and the solution is removed and cooled. Such sols contain
only a few tenths of a percent silica, but could be concentrated readily by ultrafiltra-
tion. The formation of colloidal silica in this manner was studied by Kitahara and
Oshima (78), who also examined the rate of dissolution of the particles when the sol
was diluted.

Neutralizing Soluble Silicates With Acids

When a dilute solution of sodium silicate is partially neutralized with acid to a pH of


8-9, a silica sol rather than a gel is obtained if the concentration of the resulting
sodium salt is less than about 0.3 N , and if the neutralization is carried out at ele-
vated temperature, so that the particles grow as soon as they are formed to several
millimicrons in size. Thus a 3% silica sol may be made by partially neutralizing a
dilute solution of commercial silicate with acid, according to the conditions patented
by Alexander, Iler, and Wolter (79). A sol of silica nuclei is first made by removing
sodium ion with an ion-exchange resin from a sodium silicate solution containing
2.2% SiO,, until the weight ratio of SiO, to Na,O is 85: 1. This dilute sol is then
heated at I00"C for about 10 min to form nuclei of desired size. Then dilute solu-
tions of sodium silicate and sulfuric acid are added simultaneously while the mixture
is stirred vigorously at 95°C over a period of 8 hr and the pH is maintained at about
9. The concentration of sodium ions must not exceed about 0.3 N or aggregation of
the particles will occur. Under these conditions, the silicic acid formed by the added
silicate and acid is deposited upon the silica nuclei, so that particles of 37 nm size,
for example, are obtained.
In another patent, Alexander and Iler (80) describe the isolation of particles
formed in the above process by coagulating them with a metal ion such as calcium,
washing the precipitate free from sodium salt, and then peptizing the product to a
more concentrated silica sol by removing the calcium ions by ion exchange, for
example.
I n making a sol by neutralizing a dilute solution of silicate with acid, it is essential
that the mixing be carried out so rapidly that none of the mixture remains in the pH
range of 5-6 for an appreciable time, since silicic acid gels almost instantly at this
pH. This requires that the acid and silicate be mixed with intense turbulence, with
either excess acid or excess silicate present to prevent local gelling. Even in the most
critical pH region, around neutrality, a mixing procedure patented by Armstrong
and Cummings (81) permits the formation of a uniform silicic acid solution at
332 Colloidal Silica-ConcentratedSols

concentrations up to 20% SiO,, which even under the most adverse pH conditions
does not gel for a few minutes.
A different approach to making colloidal silica by reacting sodium silicate with
acid involves making an acidic sol and precipitating the sodium salt in a strongly
acidic medium. The polysilicic acid is temporarily stable at pH 2, and if the sodium
salt of the acid used for neutralizing the silicate is sufficiently insoluble, it can be
precipitated and separated. Once the sodium salt is separated from the acid sol, the
polysilicic acid can be alkalinized to grow colloidal particles and stabilize the
product. or can presumably be used for other processes of growing silica particles to
the desired size. Thus Teicher (82) neutralized sodium silicate with an acid to
produce an acidic sol containing a miscible organic liquid, such as an alcohol, which
precipitates the salt. An earlier process by White (83) precipitates the sodium sulfate
from a sol made from sodium silicate and sulfuric acid by adding acetone. Mar-
cheguet and Gandon (84) instead use a material that will form an extremely insolu-
ble sodium salt, such as the reaction product of sulfite ion and glyoxal, the sulfite
being the acid used for neturalizing the alkali in the silicate. By reacting sodium sili-
cate with oxalic acid to obtain an acidic sol, sodium is precipitated as sodium
hydrogen oxalate, leaving about 0.13 N sodium ion in solution at 15°C (85). The
remaining salt can then be removed by ion exchange (86).

Electrodialysis
Colloidal silica has been made by various procedures involving electrodialysis
whereby sodium ions are removed from a solution of sodium silicate to produce sol.
These have been reviewed by ller (8). but in no case were stable products made.
Sanchez (87) and ller (88) patented processes of electrolyzing alkali metal silicate
solution to continuously remove alkali metal ions until a sol is obtained.
Electrodialysis may eventually replace ion exchange for making commercial sols
because alkali, oxygen, and hydrogen could be recovered and there is much less
waste water containing salts to be disposed of. I n the ller process, there are three
compartments. I n the anode compartment sulfuric acid solution is circulated around
a lead anode and in the cathode compartment where sodium hydroxide is generated,
the alkali is circulated around the steel cathode. These are located on opposite sides
of two parallel, closely spaced cation-exchange membranes between which the
process solution is rapidly circulated at 60-90°C. The process solution is a silica sol
containing about 0.05 N Na,SO, as a conducting or supporting electrolyte, A solu-
tion of sodium silicate is added to the entering process stream to raise the pH to 9.5.
The current density and rate of flow of the process stream are adjusted so as to
reduce the pH to no less than 8 as the stream leaves the cell. The liberated silica is
deposited on the silica particles which are thus grown to desired size. A 25% silica
sol of 15 nm particles can be prepared directly by this process. The electrolyte is
then removed by ion exchange, the pH adjusted for optimum sol stability, and the
sol concentrated to 30-50% silica.
There is essentially no consumption of acid except the small amount needed at the
start of each batch to neutralize a dilute solution of sodium silicate (0.5% SiO,) to
pH 9 at 60-90°C to form silica nuclei to start the process. A narrow uniform spac-
Methods of Making Sols 333

ing between the membranes is required to minimize power cost and avoid silica
deposition. Water is added to the anode compartment since it is slowly transported
to the cathode compartment, from which sodium hydroxide solution is constantly
withdrawn. Anolyte and catholyte are circulated from the corresponding electrode
compartments to separators for the removal of oxygen and hydrogen gases.
A process using a mercury cathode to remove sodium from sodium silicate is
patented by Vaquero (89). No supporting electrolyte is used so that as the alkali
becomes depleted the high electrical resistance of the sol increases the power
required. A three cell arrangement using ion-exchange membranes for making sols
by electrodialysis was discussed by Prajapati and Talpade (90). No supporting elec-
trolyte was used. The final sol pH was 2-3.
A type of electrolytic process patented by Tripp (91) is used t o dissolve an anode
of silicon metal in alcohol containing a metal salt catalyst such as copper sulfate to
produce a silica organosol.
In a study of transport of silica through membranes during electrodialysis, Boari
et al. (92) found that no transport or silica deposition occurred unless the p H was
such that H S i 0 , - ions were present. This is consistent with the observation that it is
necessary to carry out electrodialysis at less than pH 9.5 (88) in order not to deposit
silica in the membrane.

I o n Exchange

T h e pioneering work of Bird (4)by which sodium was removed from sodium silicate
by ion exchange, after which the sol was concentrated by evaporation at atmospheric
pressure, led to one of the earliest stable silica sols containing around 20% silica.
Further advances by Bechtold and Snyder (6) permitted control of particle size, and
other modifications, relating t o permissible salt concentration and optimum alkali
content, by Alexander (9) and Atkins ( I O ) led to the production of a series of
concentrated sols covering a wide range of particle sizes. Thus far, the silica sols
were made by passing relatively dilute sodium silicate solution through a bed of ion-
exchange resin to produce an acidic sol relatively free from sodium, which was then
stabilized and the particles grown as desired. An alternative approach was invented
by Wolter and Iler (93), in which the hydrogen ion-exchange resin and sodium silicate
were added to a weakly alkaline aqueous reaction medium in the pH range around
9, at elevated temperature. The silica particles grew continously under these condi-
tions where they were stabilized against aggregation, so that relatively concentrated
sols in the range 10-15% silica were produced directly. An alternative procedure for
making a relatively concentrated 12% acidic sol of polysilicic acid has been patented
by Mindick and Reven (94). whereby the silicate solution is cooled and passed
through an ion exchanger so that the concentrated sol is formed at low temperature
t o avoid gelling.
Many variations in the ion-exchange procedure have been proposed. Dirnberger
(95) showed that more concentrated sols can be made without gelling by passing the
silicate solution upward through the ion-exchange resin bed, keeping the resin in
suspension. Other variations in the ion-exchange process have also been patented
(96).
334 Colloidal Silica-Concentrated Sols

The last trace of electrolytes can be removed from silica sol produced by conven-
tional means by heating and again passing it through a hydrogen ion resin to remove
alkali liberated from the particles, according to Mindick and Reven (65).
The use of a cation-exchange resin in ammonium form permits removal of sodium
from sodium silicate solution without exposing any part of the solution to low pH
where aggregation can occur. Wolter found that a 3.25 ratio silicate solution
containing up to 6% SiO, could be passed through a column of ammonium resin
without gelling. More silicate can be added to the alkaline effluent and the solution
again treated with ammonium resin (97).
For sols of very small particle size, Shannon (98) added sodium silicate to an
acidic suspension of resin until the silica concentration reached 8%. The sol was
removed and alkalinized with N a O H and N H , to have 1 % N a and 3% NH, based
on the SiO,.
When making silicic acid by passing a solution of sodium silicate containing more
than 3-470 SiO, through a bed of hydrogen ion-exchange resin, silica gel is formed
within the pores of the conventional resin. This not only results in a loss of silica and
the need to clean the bed, but also causes the resin granules to disintegrate. Accord-
ing t o Hoffman (99) a silicic acid effluent containing up to 6% SiO, can be made
without these difficulties if one uses a macroreticular cation-exchange resin
(Amerlite IR-200) having pores about I O nm in diameter, a porosity of 3270, and a
surface area by nitrogen adsorption of 42 m z g - ’ . The resin is treated with caustic
solution to remove small amounts of silica after each use and before regeneration.
Ion exchangers can be regenerated electrolytically according to Matejka (100).
Instead of an ion-exchange resin, a cation-exchange membrane can be used to
remove sodium from a silicate solution into sulfuric acid (101). A hot sol of nuclei
particles is circulated rapidly through tubing of ion-exchange polymer which is
immersed in dilute sulfuric acid. Sodium silicate is added to the sol at a rate to
maintain the pH around 8-10 and the liberated silica is deposited on the sol parti-
cles, thus increasing their size. S o m e sulfate ions penetrate the membrane so that the
concentration of sodium sulfate in the sol slowly increases with time. The sol can be
purified and concentrated by ultrafiltration, but the sol concentration must be main-
tained so that the sodium i n normality N does not exceed N = 0.26 - 0.005C -
O.OOl2(T - 40), where C is grams of SiO, per 100 ml and T is the temperature in
degrees centigrade.

Peptizing Gels

As early as 1864, Thomas Graham reported that silica gel could be liquefied by a
trace of alkali, which he described as “peptization of the jelly.” In 1922, Praetorius
and Wolf (102) produced silica sol from a gel by heating it in water at elevated
temperature and pressure. Neundlinger (103) prepared sols containing about 10%
silica by treating the gel first with ammonia and heating without evaporating water
until a sol was produced. Improved similar processes were invented by White (5) and
Trail (104).
Methods of Making Sols 335

A modification of the preparation of sol from gel is described by Ahlberg and


Simpson (105). who formed the gel under alkaline conditions by incompletely neu-
tralizing the alkali of a silicate such as sodium silicate with less than the equivalent
acid, then washing out the salts and heating the wet gel to peptize it. Much higher
conversion of gel to sol is claimed than when an acidic gel is first made. Characteris-
tics of sols made by this process are not available, but probably sols 15-45 nm in
diameter were produced. A similar process had been described by Legal (106).
Conversion of gel to sol in an autoclave to obtain a 30% ammonia-stabilized sol was
patented by Mertz (107). The effect of ultrasonic dispersion of silica gel was
examined by Bubyreva and Bindas (108).

Hydrolysis of Silicon Compounds

In 1944, Radczewski and Richter ( 6 7 ) reported that very pure silica sols prepared by
hydrolyzing silicon tetrachloride from which the acid has been removed to give a pH
of 6.8 formed spherical particles up to 200 nm in diameter, and that these appeared
to be spongelike under the electron microscope. More recently, Stober and Fink (68)
discovered that very uniform spherical silica particles of almost any desired size up
to I micron could be made by hydrolyzing a lower alkyl silicate in an alcohol
medium containing suitable amounts of water and a m m o n i a . Subsequently,
Flachsbart and Stober (109, 1 IO) were able to incorporate certain radioactive tracers
without affecting the growth of the silica particles during their formation, but it was
not certain whether the metals were homogeneously incorporated throughout the
spheres. Particle sizes from 50 t o 900 nm in diameter were produced. It is likely that
the large spherical particles are actually spherical aggregates of much smaller parti-
cles, IO nm or less in ‘size, a s described by Radczewski.
According to Brinsmead and Brown ( I 1 I ) a “silicic acid sol” containing 43% SiO,
was made by refluxing a mixture of ethyl silicate and isopropyl alcohol to which a
dilute aqueous solution of acid was slowly added t o furnish a stoichiometric amount
of water for hydrolysis. There was no evidence that hydrolysis was complete, as
would be shown by complete miscibility with water, nor any indication that discrete
silica particles had been formed.
Sodium silicate can be hydrolyzed i f sufficiently diluted. Thus a dilute solution of
high ratio sodium silicate can be hydrolyzed in an autoclave to form colloidal parti-
cles which coagulate t o a precipitate if the sodium ion concentration is sufficiently
high. Schniirch and Koster ( I 12) report that when a solution of sodium silicate with
a 50,:Na,O weight ratio of 3.89: I is diluted to 20 g I - ’ of S O , and heated 1.5 hr
at 15OoC,38% of the silica is precipitated as fine silica.

Dissolution of Elemental Silicon

If pulverized silicon metal is first treated to remove the oxide film with hydrofluoric
acid, it reacts rapidly with water in an alkaline medium, particularly ammonia, to
336 Colloidal Silica-Concentrated Sols

form colloidal silica which is stabilized by the alkali. This type of process has been
patented by Balthis (69); the process is accelerated by carrying out the reaction while
grinding the silicon ( I 13).
Sols made from elemental silicon range from 8 to 35 nm in diameter, and under
some conditions up to 150 nm. The process and the nature of the product has been
discussed under the heading of "porous particles."
To make sols containing up to 50% SiO, by dissolving silicon, Bobb ( I 14) claims
the use of an aqueous solution of an inorganic base ( N a O H . KOH) to catalyze the
dissolution at 50-I00"C and stabilize the resulting sol. I t is remarkable that sols
made at 90-95°C were said to be very viscous but when made at 98-IOO"C were of
normal low viscosity. Particle sizes were between IS and 45 nm. Also the sols were
unusual in that they did not form hard gels when acidified. but only soft coacervates.
These differences are unexplained. An alcosol is obtained when silicon is used an an
anode and dissolved by electrolysis in an alcohol-water mixture. Acid or a metal salt
is added to provide conductivity according to Tripp (91) or Chilton ( I 15).

Dispersion of Pyrogenic Silica

Silica vaporizes only at very high temperature, around 2000"C, but if a reducing
agent is present so as to form silicon monoxide, SiO, the sublimation temperature is
1700°C. As the monoxide evaporates in an oxidizing atmosphere, the dioxide is
condensed in an extremely finely divided form. Ethyl silicate can be oxidized and the
resulting SiO, vapor condensed. The most common process involves the combustion
of silicon tetrachloride with natural gas, forming hydrogen chloride and silicon
dioxide vapor, which condenses to a very voluminous powder. By controlling com-
bustion conditions, presumably in a manner similar to that employed in making
carbon black. products of different ultimate particle sizes and degrees of coalescence
of the particles can be made. Another process involves vaporizing silica in an electric
arc and condensing the resulting vapors. Powders of this type are considered here
only because colloidal dispersions can be made from some of them. The processes
involved and the silica powders a r e discussed in Chapter 5.
Dispersion of silica to a sol of separate, discrete, ultimate particles is difficult
because the particles are coalesced to varying degrees. Also in many cases, the sur-
face is partly anhydrous with only a few hydrophilic silanol groups. For these
reasons the properties of sols of this type are generally different from those made in
aqueous solution. They d o not form strong gels and are of little use as an inorganic
binder.
The patent literature suggests that intense mechanical shearing forces, both on the
dry voluminous powder and on the subsequent suspension in water, are required to
attain reasonable dispersion in water or polar organic liquids. In water, wetting
agents are used to promote wetting of the hydrophilic siloxane surface areas and
alkali to promote surface hydration and dispersion ( 1 16, 117).
A more stable dispersion is obtained by adding t o the silicon tetrachloride a
certain amount of titanium or aluminum chloride to produce silica containing a
Purification. Concentration, Preservatives 331

small amount of the metal oxide. Very stable concentrated sols containing 40-60%
solids are obtained ( I 18). A silica produced by flame hydrolysis by Degussa, for
example, containing 1.3% aluminum oxide based on the silica (Aerosilj MOX), is
sold particularly for making concentrated aqueous dispersions with ultimate parti-
cles in the range of 20-40 nm diameter, along with many smaller ones ( I 19).
Flame-hydrolyzed silica with specific surface areas in the range of 200-400 m2
g - ’ , under the name of Cab-O-Silnqis dispersible in water at pH 9 with ammonia,
for example, to give sols up to 30% by weight of silica, provided the material is
passed through a homogenizer to break apart the three-dimensional network of ulti-
mate particles. The resulting particles still consist mainly of chainlike aggregates
which increase the viscosity (120).
In general, the “fume” or “flame hydrolysis” process does not yield silica that is
dispersible in water to give sols of discrete particles with the low viscosity a t high
concentration that is characteristic of sols made by aqueous polymerization
processes. Nevertheless, with enough processing, pyrogenic silicas of ultimate
particle size of 10-25 nm can be disaggregated and dispersed to aquasols containing
up to 40% silica with suitable mechanical treatment and dispersing agents.
Extremely finely divided silica, composed of particles only 25-50 A in size and
having a specific surface area as high a s 1000 m* g - l has been made, according to
Spencer, Smith, a n d C o s m a n ( I Z I ) , by treating moist carbon black with
dimethyldichlorosilane and then burning away the organic material in air at 500°C.
The resulting silica, amounting to 4% of the original weight of the carbon black, was
an opalescent powder in the form of small spheres, about 1 mm in diameter, of
about the same size a s the carbon black pellets. The specific surface area of the silica
was 1094 m2 g-*, indicating an ultimate particle size of 2.5 nm. The material dis-
persed easily in water, with the particles forming chains up to 20 microns long, sug-
gesting that they were partially hydrophobic.
Colloid milling of pyrogenic silica in water in the presence of boric acid or alkali
borate is disclosed by Clapsdale and Syracuse (122). A 30% sol can be prepared.
Some additional patents on making sols from pyrogenic silica mainly involve the use
of alkali stabilization with sodium silicate, sodium hydroxide, hydrazine,
hydroxylamine, or mixtures with pyrogenic metal oxides ( I 23-125).

PURIFICATION, CONCENTRATION, PRESERVATIVES

Sols made by some processes contain salts or other materials that must be reduced
or removed before the sol is finally concentrated.

Ion Exchange

Special purification procedures t o remove salts from the final concentrated sols
usually involve treatment with ion-exchange resins t o remove soluble salts and then
stabilization with a minimum of base, including ammonia, to obtain a sol of
338 Colloidal Silica-Concentrated Sols

maximum purity, according t o Rule (7). The method has become so commonplace
as to require no further discussion. Sodium is difficult t o remove after particles have
been formed; Schaefer and Gamage use an alkanolamine as the base during particle
growth (126a).

Dialysis And Electrodialysis

Dialysis is the oldest purification procedure for removing soluble impurities from
sols. Once dilute silica sol had been by reacting acid and silicate, or by hydrolyzing a
material such as silicon tetrachloride, it was early recognized that purification
required removal of the electrolyte. Graham (126b). one of the earliest investigators
of silica sols, in 1861 used dialysis to remove the electrolytes from the silica and thus
prepared relatively pure colloidal silica. Since dialysis is relatively slow, it is not
much used on an industrial scale. A more rapid process, not requiring tubing or flat
membranes, has been proposed which involves passing the sol through a column or
bed filled with a swollen polymer gel with such a fine pore structure that soluble
salts, but not colloidal particles, can penetrate. The gel may be of regenerated cellu-
lose or gelatin cross-linked with formaldehyde. The bed is regenerated by washing
out the salt (127).
More rapid purification can be effected by electrodialysis where a direct current
transports ions t o the membranes, Improvements are constantly being made in
equipment ( 128). especially for desalting seawater. However, for purifying silica sols
the process has not replaced the use o f ion-exchange columns, in which the invest-
ment is lower.

Washing Procedures

Methods that concentrate the sol particles but not the soluble components can be
used to purify the sol by rediluting the concentrated sol with pure water and repeat-
ing the process.
Sedimentation by gravity is generally t o o slow, but centrifuging permits
reasonably rapid concentration of silica particles larger than about 30-50 nm.
Alternatively. silica can be flocculated with divalent metal ions and the precipitate
washed free of soluble salts, then peptized by removing the flocculating ion.
Alexander and Iler (80) used ions such as MgZ*,C a z + , and Baz- and removed them
by ion exchange or, i n the case of Baz*, precipitation a s the insoluble sulfate.
Concentration by ultrafiltration or electrodecanting can also be used as part of the
washing procedure.

Concentration

A variety of methods for concentrating sols are available but evaporation of water
has remained the most common industrial procedure. However, with rising cost of
power and steam, other known methods will no doubt be considered.
Purification. Concentration. Preservatives 339

Evaporation of Water

For stable sols such as colloidal silica, forced circulation evaporators have been
generally used. Special precautions must be taken not to permit the sol to become
too concentrated or to reach dryness at any point on the equipment walls and espe-
cially on the heat exchange surfaces. I f this occurs, a layer of hard adherent silica is
built up. The problem becomes acute as the silica approaches the final high
concentration with increasing sol viscosity.
Evaporation has the advantage that high temperature often plays a significant role
in consolidating the structure of somwhat porous silica particles that were made at
lower temperature, and also in contributing to further growth in particle size.

Centrifugation

This is generally used for removing debris and clarifying silica sols but not for
concentrating them. Unless the particle size is greater than about 30 nm, very high
speeds are required. Hence the method is not practiced for most commercial sols.

Ultrafiltration

Remarkable advances have been made in the last 20 years in equipment and
membranes used in the process of ultrafiltration as shown in Figure 4.3. Basically,
the process removes water and small ions and solutes from a sol or colloidal suspen-
sion, which is thus concentrated without forming a filter cake or deposit on the
membrane filter.

FILTER

CAKE

ULTRAFILTER

CO~C Figure 4.3. Filter retains large particles


as a filter cake but passes colloidal parti-
cles (black) and salts. Ultrafilter retains
0 colloidal particles as a concentrated sol
- but passes soluble salts.
340 Colloidal Silica-Concentrated Sols

This technique is becoming increasingly important for purifying and concentrating


colloids with minimum consumption of energy. Thus a process described by Iler
(129) makes i t possible to make colloidal silica by partly neutralizing a hot solution
of sodium silicate with acid a t such dilution that the resulting particles are not coag-
ulated by the sodium salt. The sol (2-3% SiO,) is cooled t o 50°C and concentrated
by ultrafiltration while salt is simultaneously washed out by adding water. T o avoid
aggregation or formation of microgel, the water must be added at such a rate as to
keep the salt concentration below a certain normality N where

N = 0.26 - 0.005C - 0.002 ( T - 40)


and C is the concentration of silica in grams per 100 ml. The temperature, T‘C, is
gradually raised to 75°C while the salt normality is reduced t o below 0.15 N and C is
increased to IO. Ultrafiltration and washing are continued until a stable 30-4070 sol is
obtained. Sols of particles smaller than IO nm require that during the process, the salt
level be kept somewhat lower than with larger particles.
Reducing the p H of a silica sol to 2-4 before concentrating by ultrafiltering was
proposed by Chilton (130). However, such sols are unstable a t high concentration
and especially at high temperature where ultrafiltration is much more rapid.
Improved membranes have been the key t o recent advances in ultrafiltration. The
finest filter papers have pore diameters of as small as 1000 nm (I micron) whereas
ultrafilter membranes can be made with pore diameters from 1000 nm to as small as
2-3 nm. For many years “cellophane” or freshly formed films from collodion
(nitrocellulose) were used, but now a number of manufacturers supply strong, flexi-
ble. and durable membranes of remarkably uniform pore size yet with high porosity,
permitting rapid flow of water. Porous glass membranes have also been developed as
well as porous carbon. Porous ceramic with a microporous coating provides an
ultrafilter highly resistant to high temperature and chemical attack.
The development of membrane ultrafiltration for the chemical process industry
was described by Michaels (13 I ) , who reviewed the basic principles, equipment, and
types of applications. The useful ranges of available separation processes for
molecules and particles of different sizes were compared by Porter and Michaels
(132). Available membranes were described with pores of uniform sizes of selected
diameters over the range 1-20 nm (10-200 A). A bibliography of applications has
been compiled (133). Equipment and information a r e available from a number of
manufacturers, representing a wide range of apparatus design and membrane
construction.
A better understanding of the basic principles of ultrafiltration has permitted the
development of much more efficient equipment and corresponding commerical
applications. A detailed consideration of the theory of ultrafiltration was published
by Porter ( I 34), who dealt specifically with the problem of “concentration polariza-
tion.” As shown in Figure 4.4, the movement of particles toward the membrane
results in the formation of a concentrated sol layer of high viscosity. This can reduce
the flux or rate of filtration to a small fraction of that of the liquid medium in the
absence of colloid. The resistance t o flow is not due to plugging of pores, or even to
an actual solid layer (gel) of close-packed colloidal particles. In has been observed
Purification, Concentration. Preservatives 34 I

Figure 4.4. Ultrafiltration-concentration polarization and microgel. As dilute sol circulates


over membrane, sol is concentrated at membrane surface, and particles diffuse back to dilute
region. Microgel does not diffuse, remains at membrane.

by Iler that the resistance is a direct function of the high viscosity of the
concentrated sol layer. When water above is flushed through the system, the parti-
cles diffuse away from the boundary layer and water passes at the original flux rate.
During operation the concentration of colloid becomes “polarized” in the sense that
i t becomes more concentrated near the surface of the ultrafilter a n d less
concentrated farther away from it. This polarization is completely reversible.
To maintain a high flux the thickness and concentration of the colloid layer at the
filter surface must be minimized by creating high shearing forces in the liquid near
the surface. This is done by increasing the linear rate of flow of sol past the surface
and especially by turbulent flow. On small laboratory ultrafilters with a horizontal
membrane, high turbulence is maintained by a mechanical stirrer located very close
to the membrane. Under otherwise constant conditions, the flux rate decreases
rapidly with increasing sol concentration. In general, increasing the pressure on an
ultrafilter is less important than minimizing concentration polarization.
Flux increases rapidly with increasing temperature. Hence, where possible,
process solutions should be ultrafiltered a t maximum feasible temperature.

LIMITATIO ONN S A L TREMOVAL B Y ULTRAFILTRATION. Ultrafiltration with


continued addition of water is an efficient way t o remove salts down to a concentra-
tion of 0.03 N even from a concentrated silica sol. At these low levels there is a
tendency for the salt to remain associated with the charged surface of the colloidal
particles. This may be a hitherto unrecognized phenomenon, at least in the
chemistry of colloidal silica, but it must have been noted during removal of salt from
other ionic colloids. It is suggested that in dilute sols where the charged particles are
far apart and there is a high concentration of sodium counterions around the parti-
cles, there is a tendency for sulfate ions to be concentrated as a secondary layer out-
side of the layer of sodium ions. As shown in Figure 4.5, around each silica particle
342 Colloidal Silica-Concentrated Sols

there is a boundary layer containing a preponderance of N a + ions. Immediately out-


side of this there m u s t be a secondary layer that contains a greater number of nega-
tively charged sulfate ions than of positive sodium ions. Thus there tends to be a
higher concentration of sulfate ions in the neighborhood of the particles than in the
intervening water. so that there is less sulfate in the filtrate.
The effect becomes marked in sols of small particles containing less than 20%
silica at pH 9. By measuring the concentrations of SO,'- in the ultrafiltrate 2nd in the
sol the following typical observations were made.
To 6.5 liters of 9 5 SiO, sol of 6 nm particles at p H 9 containing 0.04 N sodium
sulfate. IO liters of wash water was added while removing filtrate to keep the sol a t
constant volume. Calculations show that this should have reduced the salt content to
0.009 ,Y. but actually i t was 0.022 A'. Simultaneous analyses verified that the
concentration of sulfate in the filtrate was less than in the sol during the operation.
In the final sol there were present 0. I 5 sulfate ions and 0.8 counter sodium ions per
square nanometer of silica surface. I t is of course possible to further reduce the
sulfate content by continued washing, but this process soon becomes very inefficient.

E F F ~ COF
T M I C R O G EILX S I L I CSOL.
A The most critical factor in using ultrafiltra-
tion to concentrate silica sols is the presence of rnicrogel or silica aggregates which
are of the order of half a micron or more in size. Aggregates this large diffuse so
slowly in comparison with single particles that they are carried by the flux to the
membrane surface where they are deposited irreversibly as gel and reduce the flow of
water and ions. In a series of ultrafiltration tests on a 4 % SiO, sol the flux was

Figure 4.5. Salt retention by colloidal particles. Around a negatively charged silica particle
there is a layer of counter sodium cations. Outside of this there is a layer in which sulfate
anions are more concentrated than in bulk solution.
Next Page

Purification. Concentration, Preservatives 343

DILUTE
SOL II

Figure 4.6. Apparatus for electrode-


cantation: anode i n dilute acid; three
porous membranes or ultrafilter type
. . bar-
riers open at top and bottom; cathode in
dilute alkali; cation permeable, anion

reduced by 50% when only 0.5% of the silica was present as microgel and by 80%
when 1.5% was present. Thus a s little as 200 ppm of silica in suspension caused an
immediate. serious reduction in flux.

Electrodecantarion

When a current is passed through a low salt, dilute silica sol, electrophoresis occurs
and the negative particles move toward the anode. By setting up a series of
membrane barriers in the sol between ion-exchange membranes to isolate the sol
from electrolyte-filled electrode compartments, it is possible to concentrate the sol
as shown, in principle, in Figure 4.6 (135). The particles concentrate against the
sides of the membranes facing the cathode, forming a dense liquid that sinks to the
bottom and is drawn off. For maximum efficiency the sol should be low in soluble
salts so that most of the current is carried by the particles and their counterions.

Preservatives

Surprisingly, microorganisms often grow in concentrated silica sols; they may


appear as floating viscous masses, green, yellow, red, or brown particles in suspen-
sion, or whitish fibrous aggregates. The presence of bacteria is often evidenced by
the smell of hydrogen sulfide and a dark color owing to precipitation of traces of
heavy metals. Algae can appear as green fibrous masses or brownish flocs.
Preservatives are often added such as formaldehyde ( 1 36) or a polyhydric alcohol
along with pentachlorophenate ( 137). The polyhydric alcohol, such as ethylene
glycol, stabilizes the sol toward freezing. Other patented preservatives are 3,5-
dimethyltetrahydro-1,3,5,2H-thiadiazine-2-thionealong with formaldehyde (138) or
25-200 ppm of sodium chlorite, NaCIOl (139).
For some unknown reason, sols stabilized with ammonia, even though a t the same
pH as those stabilized with N a O H , d o not support the growth of microorganisms.

Potrebbero piacerti anche