Sei sulla pagina 1di 7

JOURNAL OF COLLOID AND INTERFACE SCIENCE

ARTICLE NO.

185, 459465 (1997)

CS964564

SolGel Synthesis of Microporous Amorphous Silica


from Purely Inorganic Precursors
MARIA-JOSE MUNOZ-AGUADO

AND

MIGUEL GREGORKIEWITZ 1

Instituto de Ciencia de Materiales, CSIC, E-28049 Madrid-Cantoblanco, Spain


Received May 13, 1996; accepted August 19, 1996

Aqueous sodium silicate solutions and hydrochloric acid were


used to prepare silica gels without any additives. In order to establish the optimum conditions for the preparation of microporous
and monodisperse gels, the solgel reaction has been studied as
a function of pH, concentration of silicate, time, and temperature.
From 29Si NMR spectroscopy it was found that polycondensation
sets on immediately after mixing the reagents, with a continuous
increase of the fraction of 4-connected silicate tetrahedra ( Q 4 ) at
the expense of Q 1 and Q 2 . Polycondensation continues beyond the
gelation point for 1 pH 7, but not for 7 pH 10. Two
mechanisms were also observed for the gelation kinetics which is
about second order with respect to [Si] for pH 7, whereas for
pH 7, the order increases rapidly with pH. Gels formed at pH
7 had an electrostatic nature and redissolved in water, whereas
for pH 7, gelation was irreversible as expected for a gel where
individual colloid particles are linked by chemical (GSi{O{
SiG) bonds. From nitrogen adsorption, it is found that the pore
size depends mainly upon pH, the finest pores being obtained for
pH 23 (BET surface area 650700 m2 /g, porosity 35%, pore
radius 511 A). q 1997 Academic Press
Key Words: water glass; solgel synthesis; microporous silica;
inorganic membranes; gas diffusion.

INTRODUCTION

Recently, active research in ceramic membranes has been


carried out because, as compared with organic membranes,
they have important advantages which derive from the chemical and thermal stability and the high mechanical strength
of inorganic materials.
A very interesting future application of ceramic membranes is gas separation at high temperatures ( 3007C),
e.g., in membrane reactors. For gas separation purposes, it
is vital to decrease the pore size in order to obtain high
selectivities. Suitable materials are therefore microporous
(pore radius less than 1 nm) or dense, and diffusion will
be slow. Nevertheless, acceptable fluxes can be obtained if
1
To whom correspondence should be addressed. Present address: Dipartimento Scienze della Terra, Universita`, Via delle Cerchia 3, I-53100 Siena,
Italy. E-mail: gregor@unisi.it.

the microporous material is deposited as a thin layer on the


top of a porous support which consists generally of a similar
material with a coarser pore structure.
Amorphous silica, which can be obtained with one of the
highest known degrees of dispersion (700 m2 /g, (1)), is a
promising candidate to be used as a microporous top layer.
First studies in this direction have recently been reported
and show that a silica top layer could be formed either by
deposition from the gaseous phase (24) or by solgel
techniques, the latter using either organic (5, 6) or inorganic
(79) precursors for the polycondensation. However, the
resulting membranes had so far a poor definition with regard
to cracks, thickness, and pore size distribution, and almost
nothing is known about their temperature resistance. In the
solgel techniques, these problems are obviously related
to organic compounds (unavoidably present in the case of
organosiloxane precursors, and so far always employed as
additives in the case of inorganic precursors) which can
be expected to introduce important structural uncertainties
during the drying and eventual calcination of the gel.
It seems therefore interesting to try the fabrication of microporous silica membranes from purely inorganic precursors, using a silica sol obtained from sodium silicate. The
present study is concerned with the first step to this end, i.e.,
understanding the solgel reaction in order to optimize the
conditions for obtaining stable, homogeneous, and microporous gels, with a gel time which would be operative, later
on, for the fabrication of membrane top layers.
MATERIALS AND METHODS

About 500 silica gel samples were prepared by neutralization of sodium silicate solutions (Panreac, 6 mol/liter SiO2 ,
SiO2 /Na2O 3.3 by weight) with hydrochloric acid (Carlo
Erba p.a., 2 mol/liter HCl).
In order to conduct polycondensation at a defined pH, the
method of Ref. (10) was followed: hydrochloric acid and
silicate solution were added simultaneously to distilled water
in a polypropylene beaker, and the mixture was vigorously
stirred to avoid local changes in pH during the formation of
the precursor sol. The silicate solution was added through a

459

AID

JCIS 4564

6g1a$$$741

12-18-96 14:34:36

0021-9797/97 $25.00
Copyright q 1997 by Academic Press
All rights of reproduction in any form reserved.

coida

460

OZ-AGUADO AND GREGORKIEWITZ


MUN

peristaltic pump at 0.5 ml/min, and HCl was dosified using


an automatic burette (radiometer) coupled to a pH meter in
order to maintain a constant pH ( {0.1) at a previously chosen value between 1 and 10. For a final silica concentration
of 0.4 mol/liter in about 10-ml reaction volume, as an example, the required molar ratio HCl/SiO2 varied from 0.72
for sols with pH 1 to 0.43 for sols with pH 9. A glass
electrode against calomel (response time 1 s) was used
to measure pH, and standard buffer solutions of pH 4 and
7 served for calibration of the pH meter before each batch.
After preparation, the sols were transferred into closed plastic vessels and allowed to age at a controlled temperature
until gelation occurred. No change of pH was observed during this period.
During the aging period, the polycondensation reaction
was assessed by 29Si nuclear magnetic resonance spectroscopy until the solgel transition point, where the lines increase rapidly in width and become practically unobservable
at the small Si concentration in the hydrogels. A Bruker
MSL 400 instrument was used, at 79.49 MHz and pulses of
p /2 5 ms. The degree of condensation of the silicate anion
is expressed in the commonly used Q n notation (11). In this
notation, Q represents a silicon atom bound to four oxygen
atoms forming a tetrahedron, and the superscript n indicates
the connectivity, i.e., the number of other Q units attached
to the SiO4 tetrahedron under study.
The gel time (time of gelation tg ) was defined as the time
which elapsed between the preparation and the gelation of
each particular mixture. The point of gelation was taken as
the earliest moment at which the gel broke away from the
wall instead of flowing as a liquid when the beaker was
tilted (12).
The chemical reactions that cause gelation continue long
after the gel point and produce a strengthening, stiffening,
and shrinkage of the network, which results in the spontaneous expulsion of solvent from the pores so that, when working with closed vessels, a supernatant liquid will be observed. This phenomenon is called syneresis and its rate was
determined through the weight loss of the gels after removal
of the expelled solvent during a period of up to 70 days.
Such aged gel samples were not used for further experiments.
For all following determinations, freshly prepared hydrogels served as a starting point. Once gelled, these samples
were washed with distilled water during several days until
there was no reaction for Cl 0 in the washing water with
respect to Ag / , and finally, they were rinsed, filtered, and
allowed to dry at room temperature in order to obtain the
corresponding xerogels. This step took about 1 month and
led to transparent monolithic blocks which were crushed to
obtain a powder.
The surface area, pore volume, and average pore radius
in the xerogels were determined by nitrogen adsorption
desorption isotherms after dehydration at 1707C for 18 h in
vacuo. The measurements were carried out at 01977C in

AID

JCIS 4564

6g1a$$$742

12-18-96 14:34:36

a Micromeritis Asap 2000. The specific surface area was


calculated by the BET method (13), the pore volume from
extrapolation of the isotherm until a relative pressure of 0.94,
and the mean pore radii from the values for the cumulative
pore volume and surface as obtained from the desorption
branch of the isotherm using the BJH (14) method.
For chemical analysis, the xerogels were digested in
HF / HNO3 following the procedure of Bernas ( 15 ) , and
Na, Si, Cl were determined using standard analytical techniques. The water content of the xerogels was determined
by DTA / TG in a Stanton 750 heating at 10 7 /min from 25
to 9007C in static air.
RESULTS

Using the above-mentioned procedure, sols were prepared


at different pH from 1 to 10 and silica concentrations ranging
from 0.07 to 0.9 mol/liter Si as referred to the final solutions.
At lower silica concentrations, gelation did not occur during
the time period studied ( 70 days), and for higher concentrations, reactions became so fast that formation of a homogeneous gel was no longer warranted under our experimental
conditions.
For sols prepared at pH 1, 3, and 9, 29Si NMR spectra have
been registered at different stages of the polycondensation
reaction from its beginning after the mixing of reactants until
the gel formation occurred. A typical sequence is given in
Fig. 1, and the quantitative evaluation of the spectra for all
pH values is resumed in Table 1. As can be seen, condensation sets on immediately with an important increase of Q 4
at the expense of Q 2 , Q 1 , and Q 0 which is common for all
pHs. After this, acid and basic systems behave differently:
for pH 1 and 3, Q 2 / Q 1 continue to lower steadily until
the gel point, which occurs at 5153% of Q 4 , whereas for
pH 9, significant amounts of Q 2 / Q 1 exist throughout the
polycondensation which proceeds, in this case, at the expense of Q 3 until it stops at 91% Q 4 . In the solid samples,
after gelation only very badly resolved NMR spectra could
be obtained due to dipoledipole interactions and low 29Si
concentration ( 0.047 1 0.9 1 6 1 10 20 2.5 1 10 19
nuclei/cm3 ). In any case, they suggest that condensation
continues for a considerable time after gelation in the acid,
but not in the basic gels.
Another important kinetic parameter is the gel time tg ,
which is inversely proportional to the rate of gel formation.
In Fig. 2, tg-values have been plotted against the silica concentration for different pH, and in Fig. 3, the same observations have been represented as a function of pH, with the
silica concentration as parameter.
As can be seen from these figures, the gelation is faster
for high Si concentrations (Fig. 2) and also depends strongly
on pH, with a minimum gel time at pH 7 (Fig. 3). In
order to interpret these data, one may consider the rate of
gel formation as a measure of the rate of polycondensation.

coida

461

MICROPOROUS SILICA

condensation reaction follows the same mechanism which


is approximately second order with respect to Si. From the
vertical offset between the lines, as well as from Fig. 3, it
can be further deduced that the reaction is proportional to
the OH 0 concentration, so that the full mechanism could be
described by the reaction sequence
fast

GSi{OH / OH 0 r GSi{O 0 / H2O

[1]

slow

GSi{OH / 0O{SiG r

GSi{O{SiG / OH 0

FIG. 1. 29Si NMR spectra showing the polycondensation reaction in a


silica sol (pH 1, [Si] 0.9 mol/liter). The spectra were taken at different
moments t after preparation and show a progressive increase of the proportion of highly condensed SiO4 tetrahedra (Q 4 and Q 3 ). The last spectrum
(t 51 h) was taken immediately after gelation, and the first one ( t 0
h) is given for comparison and refers to the starting solution ([Si] 6
mol/liter, pH 11). Chemical shifts d are given with respect to tetramethylsilane, neighboring spectra are displaced along the ordinate by an arbitrary
amount. Q n assignments were made according to (20). The small drift
toward higher fields for corresponding configurations in the starting solution
is a common observation (see e.g., (21)) for increased pH and silica concentration.

If the rate of the condensation reaction is expressed in terms


of the conversion of SiOH to SiOSi, the rate law is
d[SiOSi]/dt } k[SiO2 ] n , and tg , i.e., the time needed to
produce the degree of conversion required to transform the
sol into a gel, directly shows the polycondensation mecha} d[SiOSi]/dt. With this assumption,
nism putting t 01
g
the constant slope (about 02.5) of the lines for 1 pH
6 (Fig. 2) indicates that, in the whole acid regime, the

AID

JCIS 4564

6g1a$$$743

12-18-96 14:34:36

[2]

which combines OH 0 catalysis Eq. [1] with a bimolecular


condensation reaction Eq. [2] as the rate-determining step.
For pH 7, this mechanism is no longer valid. On the
contrary, gel formation becomes now slower with increasing
pH (Fig. 3), and from Fig. 2 it appears, that the order with
respect to Si increases rapidly with pH, the slope of the lines
changing from 03.4 to 05.4 and 013.4 for pH 7, 8, and 9,
respectively.
It is not reasonable to assume that such high orders reflect
in any way a condensation reaction in the sense of Eq. [2],
so that an alternative mechanism for the gel formation must
be looked for. One indication comes from observations about
the chemical nature of the gels: in the acid regime (1 pH
6), the gels formed irreversibly and were stable against
dilution with water, whereas for pH 7, they became increasingly unstable until pH 9, where they dissolved entirely
in water. This suggests that, for 1 pH 6, gel formation
is based on the establishment of chemical bonds between
neighboring particles, whereas for pH 9, the gel has an
entirely electrostatic nature where particles are held in their
equilibrium positions by repulsive rather than attractive Coulomb forces.
Further information about the gelation mechanism can be
taken from the temperature dependence of the gel time (Fig.
4): in the acid milieu (pH 1, 3, and 5), gelation is seen to
be an activated process as expected for the formation of a
covalent bond in the sense of Eq. [2], whereas for pH 9,
the gel time is almost independent of temperature.
After gelation, most gels showed a syneresis which has
been measured and plotted against time during the first 70
days of aging (Fig. 5). The plots can be divided into two
regions: the first one (t 18 days) shows a deceleratory
curve and weight losses which depend strongly on pH,
whereas the second one (t 18 days) is a straight line with
similar slope for all pHs. With this behavior, the second part
is likely to reflect a simple drying process due to some
unavoidable evaporation to be separated from the proper
syneresis (spontaneous expulsion of water from the gel)
which is restricted to the first region (t 18 days) in Fig.
5. The quantity of water expelled near the end of this region

coida

462

OZ-AGUADO AND GREGORKIEWITZ


MUN

TABLE 1
Percentage of Different Species of Silicate Tetrahedra, Qn, and Its Variation with the Reaction Time, t,
for Silica Sols Prepared at pH 1, 3, and 9 and [Si] 0.9, 0.7, and 0.3 mol/liter, Respectively
pH

12

t(h)

Q0
Q1
Q2
Q3
Q4
Q3/Q4

4
10
33
45
8
5.6

1
1

3
24
50
23

16
50
34

2.17

24

14
52
34

1.47

1.53

3
31

12
49
39

51

5
42
53

1.25

0.79

17
48
35
1.37

24

8
51
41
1.24

9
47

9
46
45
1.00

72

8
41
51

5
34
42
19

0.80

Note. The percentages have been obtained from the integrated area of the corresponding peaks in the
for comparison (as in Fig. 1), and gelation falls, in all cases, between the last two points.

2.21

24

4
41
55
0.74

48

12
13
37
38
0.97

72

99

14
43
43
1.00

9
91
0.10

29

Si NMR spectra. Values for t 0 are given

has been plotted against pH in Fig. 6, which contains also


the gel time for comparison. Clearly, the weight loss by
syneresis is maximum for the same pH were gelation is
fastest. It can therefore be assumed, that the syneresis involves the polymerization of oligomers in the hydrous interstices of the gel, where they remained trapped without reacting because of the rapid gelation. This hypothesis would
be compatible with the deceleratory behavior of the syneresis
curves in Fig. 5, which is typical for a diffusion-limited
process or a reaction depending on the concentration of a
vanishing reactant.
Once the reaction mechanism had been characterized, it
was important to establish a correlation with the pore size
and homogeneity of the final product. To this end, nitrogen
adsorption isotherms have been performed for xerogels obtained under different reaction conditions, paying particular
attention to the influence of the pH during preparation of
the sol. For 1 pH 5, isotherms were of type I (for the
nomenclature see (16)), which is characteristic for micropo (17)), whereas for pH 6, a
rous materials (r 10 A
hysteresis loop between the adsorption and desorption
branch develops, indicating that capillary condensation and
) become important. Examples for
mesopores (r 10 A
both types of isotherms are given in an earlier publication

(Fig. 3 in (18)). The corresponding results in terms of surface area, porosity, and pore radius of the different samples
are resumed in Table 2 which is divided into four parts in
order to examine the influence of the different parameters
during the polycondensation reaction.
The most important parameter is observed to be the pH:
as can be seen in Table 2 (part a), the specific surface area
is practically constant at a very high value of 650700 m2 /
g for acid-prepared gels, but it decreases steadily for pH
6. The pore volume, on the other hand, shows a maximum
at 6 pH 7 and it depends less on pH, which makes that
the pore size, as calculated from 2Vp /SBET , increases with
.
pH from r 10 until 49 A
Furthermore, it is observed (Table 2, part b) that, for pH
3, the silica concentration has little influence upon SBET ,
whereas the pore radius increases slightly with [Si]. More
important is the temperature (Table 2, part c) which causes
!) and
a considerable increase in pore size (from 5 to 15 A
3
porosity (0.150.50 cm /g) when going from 4 to 707C.
Finally, aging (Table 2, part d) is observed to cause a slight
increase of the pore size which goes from r 10.7 to r
for pH 3.
15.4 A
The pore size distribution was analyzed for gels prepared at different pH and showed an essentially monodisperse structure, although in the mesoporous samples ( pH

FIG. 2. Gel time tg versus silica concentration [SiO2 ] in sols of different


pH values between 1 and 9.

FIG. 3. Gel time tg versus pH for sols of different silica concentrations


between 0.1 and 0.45 mol/liter.

AID

JCIS 4564

6g1a$$$743

12-18-96 14:34:36

coida

MICROPOROUS SILICA

FIG. 4. Gel time tg versus temperature for silica soles prepared at pH


1, 3, 5, and 9. The silica concentration is [Si] 0.7, 0.55, 0.39, and 0.34
mol/liter, respectively.

6 ) , a shoulder toward smaller pore sizes is important


( Fig. 4 in ( 18 ) ) .
The chemical composition was determined for two representative xerogel samples. The xerogel obtained at pH 3 is
almost pure silica (0.1% Na), whereas at pH 7, an important
Na-content of 7.6% was found. Considering that the Clcontent is negligible in both cases, sodium evidently substitutes hydrogen at pH 7 in the sense of an acidbase reaction
GSi{OH r GSi{ONa, may be near the surface of the
silica particles, and the presence of NaCl as an impurity can
be excluded. The H2O-content corresponding to the weight
loss in the temperature range 1004007C was similar (5 and
3.5%, respectively, for xerogels prepared at pH 3 and 7).
DISCUSSION

One issue of the present results is to give a practical basis


for the preparation of homogeneous microporous silica. As
seen from Table 2, the pore radius can be triggered between
1 and 5 nm simply by the choice of a suitable pH between
1 and 9, so that it is possible to fabricate asymmetric membrane structures by successive deposition of different gels.
The transport properties of the membrane will mainly be
determined by the microporous upper layers which form at
pH 6. For these gels it is very important to remember that,
on acidification at the beginning of the polycondensation, the
originally strongly alkaline (pH 12) silicate solution must

FIG. 5. Syneresis behavior of silica gels of [Si] 0.2 mol/liter prepared at different pH values. The time of aging is measured as beginning
from the moment of gelation.

AID

JCIS 4564

6g1a$$$743

12-18-96 14:34:36

463

FIG. 6. Effect of pH on ( l ) gel time and ( / ) weight loss through


syneresis. [Si] 0.2 mol/liter, weight loss as observed after 8 days of
aging.

pass through a region of maximum instability around pH 7


where gelation is very fast (Fig. 3). Since the original silicate solution is relatively concentrated before dilution with
HCl and H2O (see experimental procedure), Si-concentrations above 1 mol/liter may locally occur in the reaction
vessel. Such concentrations fall on the right end of Fig. 2,
and from an extrapolation of the curves, gel times of a few
seconds or even less would be expected for pH 6. Vigorous stirring is therefore of vital importance to warrant a welldefined pH for the polycondensation reaction.
Another interesting result refers to the mechanism of the
polycondensation reaction and gelation and its change with
pH. As pointed out above, for pH 7 the mechanism is
OH-catalyzed and second order with respect to [Si] which
defines the constant slope of the lines for low pH in Fig. 2.
Figs. 3 and 6 have a minimum/maximum indicating that at
pH 7, the bimolecular, OH-catalyzed condensation mechanism becomes less important than other phenomena. With
pure silica, the minimum in tg appears at pH 5 (Fig. 3.2,
p. 177 in Iler (1)) as expected from the first dissociation
constant (pKa 10 (1)) of silicic acid. The observed displacement toward pH 7 in the presence of salt (NaCl)
agrees with earlier findings (1) and has been explained by
adsorption of sodium ions on the silica surface. In this way,
the negative charges created by the progressive loss of silanol protons toward higher pH would be partly compensated
so as to delay electrostatic repulsion between particles,
which means an increase in tg (see Fig. 3), toward a more
basic milieu. The elemental analysis (see above) shows indeed that xerogels prepared at pH 7 contain important quantities of Na (not NaCl; remember that the gels have been
washed with distilled water), whereas samples prepared under acid conditions show only negligible Na-contents. The
association of alkali ions with silica has recently been reported (19) also for solgel reactions using TEOS as a
precursor.
Differences with respect to earlier studies refer to the acid
and basic ends of our curves in Figs. 3 and 6. In particular,
we never found a lowering of tg at pH 3 as discussed,
e.g., in Iler (1), but observed a steady decrease approximately first order with [OH 0 ] throughout the entire range

coida

464

OZ-AGUADO AND GREGORKIEWITZ


MUN

TABLE 2
Specific Surface Area SBET , Pore Volume Vp , and Pore Radius
r for Silica Xerogels and Their Dependence on Different Reaction
Parameters
a. Effect of pH during preparation of the sol
([Si] 0.41 mol/liter, T 207C)
pH

SBET (m2/g)

Vp (cm3/g)

)
r 2Vp/SBET (A

1
2
3
4
5
6
7
8
9

693
683
654
634
699
568
467
398
205

0.39
0.35
0.35
0.38
0.44
0.62
0.64
0.56
0.50

11.2
10.2
10.7
12.0
12.6
21.8
27.4
28.1
48.8

b. Effect of the overall silica concentration in the sol


(pH 3, T 207C)
[SiO2] (mol/liter)

SBET (m2/g)

Vp (cm3/g)

)
r 2Vp/SBET (A

0.26
0.41
0.50
0.55
0.67
0.77

644
654
776
654
716
652

0.42
0.35
0.55
0.52
0.53
0.54

13.0
10.7
14.2
15.9
14.8
16.6

c. Effect of the temperature during gelation


([Si] 0.55 mol/liter, pH 3)
T (7C)

SBET (m2/g)

Vp (cm3/g)

)
r 2Vp/SBET (A

4
20
70

598
656
674

0.15
0.36
0.50

5.0
10.9
14.8

d. Influence of aging (20 days) as seen for gels prepared at pH 3 and 7.


(T 207C, [Si] 0.26 mol/liter (0.41 mol/liter for nonaged gels))
pH

Aged

SBET (m2/g)

Vp (cm3/g)

)
r 2Vp/SBET (A

3
3
7
7

0
/
0
/

654
671
467
597

0.35
0.52
0.64
1.03

10.7
15.4
27.4
34.5

Note. All data refer to xerogels obtained from freshly prepared hydrogels
without aging, except for pH 9, where some aging was necessary to keep
the gel from dissolving in the washing step and, of course, the aged samples
cited in part d.

from pH 1 to 7. On the other hand, in the range from pH 7


to 10 (Figs. 3 and 6), the increase of tg with pH is exponential. The increase of tg toward the basic end therefore has
a different origin; i.e., it cannot be explained by a linear
relationship of tg with pH in the sense of a H / or OH 0
catalyzed reaction as for pH 7, but instead it is more likely

AID

JCIS 4564

6g1a$$$743

12-18-96 14:34:36

to reflect the increase of electrostatic repulsion between more


and more negatively charged particles.
Parallel to the minimum in tg , Fig. 6 shows a maximum
for the weight loss through syneresis at pH 7. Developing
the above ideas, the small syneresis effects observed at the
acid and basic ends again have a different origin. Toward
the acid end, the reaction becomes extremely slow and there
is enough time for an orderly polycondensation without leaving monomers trapped in interstices, so that syneresis should
not be important. Toward the basic end, instead, syneresis
is small because water becomes an integral constituent of
the gel, which is now physical in nature and formed by
immobilization of charged particles. A good case for this
model was the behavior of the fresh alkaline hydrogels
against water: They peptize after adding a small quantity of
water, but after some time for repose, regelation occurs with
a slightly higher structural H2O content.
A last evidence for a completely different reaction mechanism toward the basic end comes from the apparent increase
in reaction order with respect to Si at high pH, as seen in
Fig. 2 and discussed in an earlier section. We are not able
to give a definite explanation of these findings in terms of
a mechanistic model, but tentatively we think that the increase in solubility of SiO2 toward the basic pH might play
an important role. New experiments covering a broader
range of silica concentrations and reaction times up to one
year or more would be needed to resolve this interesting
problem.
SUMMARY

The present results show that simple Na-silicate solutions


can be used for the preparation of well-defined, purely inorganic and microporous silica. The product is free from NaCl
crystals which would give rise to inhomogeneities and voids
formed during the preparation of membranes, and pore radii
by the choice of a
can be tuned from about 10 to 50 A
suitable pH during polycondensation. Slightly smaller pore
can be obtained at lower temperaradii down to about 5 A
tures, whereas aging of the hydrogels increases the pore size.
Gel times depend on the silica concentration and lie, for 0.2
0.5 mol/liter, in the desired range for practical applications
(minutesdays). In addition to the technologically important results, the present study revealed an interesting
change of reaction mechanism when passing from the acid
to the basic regime, details for the latter still awaiting an
explanation.
ACKNOWLEDGMENTS
We thank Dr. Isabel Sobrados for realization of the NMR spectra. Financial support from CICYT, Madrid (MAT88-166) and a grant (to M.A.)
from Plan Nacional del Ministerio de Educacion y Ciencia, Madrid, are
gratefully acknowledged.

coida

MICROPOROUS SILICA

REFERENCES
1. Iler, R. K., The chemistry of silica. Wiley, New York, 1979.
2. Gavalas, G. R., Megiris, C., and Nam, S. V., Chem. Eng. Sci. 44, 1829
(1989).
3. Okubo, T., and Inoue, H., J. Membr. Sci. 42, 109 (1989).
4. Li, D., and Hwang, S-T., J. Membr. Sci. 59, 331 (1991).
5. Uhlhorn, R. J. R., Keizer, K., and Burggraaf, A. J., J. Membr. Sci. 66,
271 (1992).
6. Brinker, C. J., Ward, T. L., Sehgal, R., Raman, N. K., Hietala, S. L.,
Smith, D. M., Hua, D. W., and Headley, T. J., J. Membr. Sci. 77, 165
(1993).
7. Asaeda, M., and Du, L. D., J. Chem. Eng. Jpn. 19, 72 (1986).
8. Asaeda, M., Du, L. D., and Fuji, M., J. Chem. Eng. Jpn. 19, 84 (1986).
9. Larbot, A., Julbe, A., Guizard, C., and Cot, L., J. Membr. Sci. 44, 289
(1989).
10. Okkerse, C., and de Boer, J. H., J. Chim. Phys. 57, 534 (1960).
11. Engelhardt, G., Jancke, H., Hoebbel, D., and Wieker, W., Z. Chemie
14, 109 (1974).

AID

JCIS 4564

6g1a$$$744

12-18-96 14:34:36

465

12. Merrill, R. C., and Spencer, R. W., J. Phys. Chem. 54, 806 (1950).
13. Brunauer, S., Emmett, P. H., and Teller, E., J. Am. Chem. Soc. 60, 309
(1938).
14. Barrett, E. P., Joyner, L. G., and Halenda, P. P., J. Am. Chem. Soc. 73,
373 (1951).
15. Bernas, B., Anal. Chem. 40, 1682 (1968).
16. Brunauer, S., Deming, L. S., Deming, W. S., and Teller, E., J. Am.
Chem. Soc. 62, 1723 (1940).
17. Sing, K. S. W., Everett, D. H., Haul, R. A. W., Moscou, L., Pierotti,
R. A., Rouquerol, J., and Siemieniewska, T., Pure Appl. Chem. 57, 603
(1985).
18. Munoz-Aguado, M. J., Gregorkiewitz, M., and Bermejo, J., J. NonCryst. Solids 189, 90 (1995).
19. Sanchez, J., and McCormick, A., Chem. Mater. 3, 320 (1991).
20. Engelhardt, G., and Michel, D., High-Resolution Solid-State NMR
of Silicates and Zeolites. Wiley, Chichester, 1987.
21. Harris, R. K., and Knight, C. T. G., J. Chem. Soc. Faraday Trans. 2
79, 1525 (1983).

coida

Potrebbero piacerti anche