You are on page 1of 8

]OURNA L OY

ELSEVIER Journal of Non-CrystallineSolids 186 (1995) 96-103

Preparation and properties of monolithic silica xerogels


from TEOS-based alcogels aged in silane solutions
S. H~ereid, M. Dahle, S. Lima, M.-A. Einarsrud *
Department of lnorganic Chemistry, Norwegian Institute of Technology, University of Trondheim, 7034 Trondheim, Norway

Abstract

In previous work, tetramethoxysilane-based alcogels were strengthened and stiffened by providing new monomers to the
alcogel subsequent to gelation. It is shown that also by aging tetraethoxysilane (TEOS)-based alcogels in solutions of
TEOS/ethanol (EtOH) the strength and stiffness can favourably be increased and hence reduce the shrinkage during drying
at ambient pressure. Low-density silica xerogels ( p --- 0.24 g/cm 3) are obtained by changing aging parameters such as time
and temperature. The water content in the pore liquid is important for the aging process and is increased by changing the
pore liquid prior to providing the monomers for aging. The shear modulus has been increased by 23 times and the modulus
of rupture by five times, reducing the corresponding shrinkage during drying to --- 0.5%.

I. Introduction such as strength, stiffness and permeability might,


however, be influenced by aging of the wet gel
For a large-scale commercial application of aero- [6-11].
gels, ways to circumvent the high costs and risk of Our previous work has been concentrated on in-
supercritical drying must be found. In recent years creasing the strength and stiffness of the wet gel by
processes for producing monolithic aerogels using aging the gel in a silane solution [12]. New monomers
ambient pressure drying have been described [1,2]. are added to the already formed network and hydrol-
The achievement of this goal depends on factors ysis, condensation and specific precipitation of these
such as the pressure gradient developed during dry- monomers favorably increase the strength and stiff-
ing and the strength and permeability of the wet gel. ness of the alcogel. For tetramethoxysilane
Because of the higher permeability of base-catalyzed (TMOS)-based gels, the shear modulus, G, was in-
gels compared with acid-catalyzed gels, they are creased by 15 times and the corresponding shrinkage
favourable as precursors for ambient pressure dried was reduced by about 30% giving xerogels with a
aerogels [3]. However, the strength and stiffness of considerably lower density compared with unaged
these base-catalyzed gels are generally lower than samples [12,13].
for acid-catalyzed gels [4,5]. Properties of the gel The present work examined the effect of aging
tetrathoxysilane (TEOS)-based alcogels in solutions
of TEOS/ethanol. The increase in strength and stiff-
* Corresponding author. Tel: +47-7 359 4002. Telefax: +47-7 ness of the wet gels has been reported as a function
359 3992. E-maih einrud@kjemi.unit.no. of aging time and temperature in the TEOS/ethanol

0022-3093/95/$09.50 © 1995 Elsevier Science B.V. All rights reserved


SSDI 0022-3093(95)00039-9
S. Hcereid et al. /Journal of Non-Crystalline Solids 186 (1995) 96-103 97

solution. Properties of the corresponding ambient before measurements of surface area, density, linear
pressure dried xerogels are compared with CO 2 su- shrinkage and skeleton density, and before gravimet-
percritical dried aerogels. ric measurements.
The mechanism for the aging process will be The load deflection data for the measurement of
discussed. To show the importance of the specific the wet gel shear modulus, G, and the modulus of
precipitation of the TEOS monomers from the aging ruptures (MOR) were determined by using a three-
solution, properties of these gels have been com- point beam bending apparatus (ASTM C674-81).
pared with thermally aged alcogels with the same The solution in the bath with the bending fixture was
silica content prepared from a sol with a decreased the same as the pore liquid, which was ethanol in all
amount of solvent. cases except for gels only washed in 20 vol.%
H 2 0 / E t O H at 60°C. The ratio between the span of
the bending fixture and the wet gel radius was = 10
2. Experimental procedure and the deflection versus load was measured at a
down drive rate of 2.2 mm/min. The MOR was
Alcogels were made from TEOS, H20, ethanol, obtained from the load at which the rods broke and
HC1, and NH4OH in the total molar ratio the G modulus was obtained from the slope of the
1 : 3.5 : 3.9 : 7.8 × 10 - 4 : 5 . 7 X 10 - 3 by a similar load-deflection curve. Nine rods were broken to
procedure as described by Brinker et al. [14]. The obtain an average value for the MOR. Load-relaxa-
TEOS was hydrolyzed in a solution with molar ratios tion experiments were done as described by Scherer
T E O S / H 2 0 / e t h a n o l / H C 1 equal to 1 : 1 : 3.9 : 7.8 × [15] to measure the permeability, D, the G modulus
10 - 4 under reflux for 1.5 h at 60°C. This solution and the Young's modulus, E. The relaxation time
was mixed with the remaining water and NHaOH , was obtained by measuring the load necessary to
cast into Teflon tubes with an inner diameter of 8.6 keep the deflection of the gel rod constant. The
mm and kept at 40°C for 1 h for gelation. To study deflection of the gel rod was chosen to give an
the effect of aging time and temperature in acceptable load ( = 30 g) and the data analysis was
TEOS/EtOH, the gel rods were soaked first in a performed using a non-linear fitting routine de-
washing solution (20 vol.% H20/ethanol) for 24 h scribed by Scherer [16].
at 60°C and secondly in an aging solution (70 vol.% The surface area of the supercritically dried aero-
TEOS/ethanol) for various lengths of time (6-96 h) gels was measured by N 2 adsorption (Micromeritics,
and at various temperatures (40-70°C). The samples ASAP 2000) and calculated from five-point
were then washed in ethanol four times within 24 h Brunauer, Emmett and Teller (BET) theory and the
at 50°C before the mechanical properties were mea- skeletal density was measured by helium pycnometry
sured and further in acetone four times within 24 h at (Micromeritics, Accupync 1330). The samples were
room temperature. As a reference sample for 0 h heat treated at 250°C under vacuum for 24 h prior to
aging time, gels were only washed in 20 vol.% surface area measurements. The xerogel and aerogel
H 2 0 / E t O H . The gels were soaked in the washing bulk densities were calculated from the weight and
solution for 24 h at 60°C and the mechanical proper- volume of the samples. The wet densities are the
ties were measured prior to further washing in ethanol bulk densities of dried gels if there were no shrink-
and acetone. Samples were also immersed directly age during drying; it is calculated from the volume
into ethanol after gelation for 1 h at 40°C or after a of the wet gel and gravimetric measurements of the
thermal aging in the mother liquor in well sealed heat-treated gel. The linear shrinkage during drying
tubes for 24 h at 60°C. Generally the volume ratio is calculated from the diameter of the wet gel rod
gel : solution was 1 : 3.1. after aging and after heat treatment.
The wet gels were dried (1) from ethanol or
acetone at ambient pressure in the temperature range
3. Results
20-180°C within 96 h, or (2) above the critical point
of CO 2 after replacing the acetone with liquid CO 2. Aging the wet gel in solutions of TEOS/EtOH
All dried gels were heat treated in air at 300°C causes silica to precipitate from the aging solution
98 S. Haereid et al. /Journal of Non-Crystalline Solids 186 (1995) 96-103

onto the silica network [13]. The effect of aging time 25 i i i I i , i I F i f i i , } , , ,

and temperature in the TEOS solution can be seen in 40°C


50°C
Figs. 1 and 2, where the physical properties of 20
- - - - O - - 60°C
samples aged for different times and temperatures Z 60"C (rel.)
are compared. Aging time is the time the gel rods 15 - - - - O - - 70°C
4 70"C (rel.)
were immersed in the T E O S / E t O H solution after
washing in H 2 0 / E t O H . Fig. 1 shows the wet den- 1o
sity, the xerogel density for gels dried from acetone,
and surface area versus aging time and temperature.
5
Wet densities represent the amount of silica precipi-

0.30 40ocl /o
0.35 + 501c I /~
"~ 0.30
0.25
0.25
.~ 0.20 e¢
© 0.20
.g
0.15
~0.15
0.10 ]
0.05 ! , ' ' I , , , I , , , i , , , I,, ,"
0 20 40 60 80 100
0.50 Aging time (hours)
0.45 Fig. 2. Effect of aging time at different temperatures in the
TEOS/EtOH solution on G modulus from both standard three-
~ 0.40 point beam bending and load-relaxation (rel.) experiments and
MOR. The errors (determinedfrom standard deviation)are +5%
-~ 0.35
= in the G values and _+20% in the MOR values. Lines are drawn
0.30 as guide to the eye.

0.25

i ,,litl~l~,,lll~L,Ir-
tated from the aging solution, and is therefore a
measure of the effect of the aging process. As can be
seen from Fig. 1, the wet density increases while the
xerogel density and the surface area decrease with
increasing aging time and aging temperature. The
surface areas were measured on supercritically dried
aerogels; however no difference in surface areas was
observed between aerogels and xerogels from the
same sol and with the same aging treatment of the
650 : , , , I [ ,_1 , I ivJ
wet gel. Fig. 2 shows the increase in G modulus and
0 20 40 60 80 100
Aging time (hours) MOR with increasing aging time and aging tempera-
ture. Generally, MOR data show considerable scatter
Fig. 1. Effect of aging time at different temperatures in the since the strength is controlled by random surface
TEOS/EtOH solutionon wet density, xerogel density when dried
from acetone, and on surface area measured on supereritically flaws. The constant MOR values between 48 and 72
dried aerogels. The error is _+0.01g/cm 3 in the density and + 10 h for the samples aged at 40 and 50°C are expected
m2/g in surface area. Lines are drawn as guides to the eye. to be an experimental error and that a similar in-
S. Ha•reid et al. / J o u r n a l of Non-Crystalline Solids 186 (1995) 96-I03 99

Table 1 xerogels dried from acetone and ethanol and super-


Relaxation time % G modulus, Young's E modulus, E, and critically dried aerogels versus wet density. The den-
permeability, D, given from load-relaxation experiments
sity of the xerogels decreases with increasing wet gel
Aging time in Pwet 7" G E D density until it reaches a minimum of -- 0.25 g / c m 3
TEOS/EtOH ( g / c m 3) (s) (MPa) (MPa) (nm 2)
at 70°C (h)
in the wet density range 0.21-0.26 g / c m 3 where it
increases again. The linear shrinkage during drying
24 0.171 101 5.41 13.7 11.9
48 0.259 65.9 9.22 23.9 9.94 at ambient pressure reaches a minimum level of
96 0.281 34.6 23.5 61.9 6.93 --- 0.5% when the wet density is 0.26 g / c m 3. Gener-
ally monolithic rods were obtained. The density of
the supercritically dried aerogels increases with in-
crease in the MOR values as observed for the 60 and creasing wet density as expected since the linear
70°C samples is more probable. shrinkage during the supercritical drying is low (Fig.
Table 1 shows the effect of aging time in the 3). There is also a small decrease in the linear
TEOS solution at 70°C on relaxation time, r, G shrinkage during the supercritical drying as the wet
modulus, Young's modulus, E, and the permeability, density increases. Fig. 4 shows the surface area and
D, measured from the load-relaxation-time experi- the hydraulic radius, r h, versus wet density. The
ments. hydraulic radius was calculated from
Fig. 3 shows the density and linear shrinkage of 2Vp 2(1-p .... /Ps)

rh = SA P .... SA ' (1)

0.6 , ,,, i I , r , l , , , , [ ~ , T

1100 i , , , l l l , l l l , l l TII
0.5 ~ • Acetone
O Ethanol
~ © CO 2 1000
0.4
E 900 f
'~ 0.3 o oeO- • 00
o o,Oo ~D
800
0.2 @

7OO
,°,,p,,,ll,,i,f,,,,
, , , , I , , , , i , , J ,
40 O • Acetone - .... r .... j .... II

¢~ ~ O ~ O © Ethanol
~a 30 o co 2
o~-~ tO0
• •
•~ 20
80

.= l0 • o

l
o • 60
'-' o °S°ooocoO OO •
4O

0.! 0.15 0.2 0.25 0.3


Wet density (g/cm 3) 20
0.10 0.15 0.20 0.25 0.30
Fig. 3. Change in density and linear shrinkage versus wet density
Wet density (g/cm 3)
of xerogels dried from acetone and ethanol and supercritical dried
aerogels. The error is + 0.01 g / c m 3 in the density and _+0.5% in Fig. 4. Surface area and hydraulic radius of the xerogels versus
linear shrinkage. wet density.
100 S. Hcereid et al. /Journal of Non-Crystalline Solids 186 (1995) 96-103

- - y = 2558.7 * x^(3.3802) R= 0.98362


100 35
i(a) i6 ..... ' ......... ' ......... ' ..........
30 A 40°C
<> 50°C
25 [] 60"C
o 70°C
2o -•
l0
40.C M= 15
50"C
10
60"C
70"C 5
[]
1 0
0.1 0.2 0.3 0.4 0.5 1.0 2.0 3.0 4.0 5.0
Wet density (g/cm3) Pxero/Pwet

Fig. 5. (a) Bulk modulus versus wet density on a log-log scale and (b) bulk modulus versus the (xerogel density)/(wet density) ratio for
samples aged in T E O S / E t O H solutions for different times and at different temperatures.

where SA is the surface area and Vp is the specific to drying [1,12,13]. The data from this study give the
pore volume of the xerogel given by the xerogel bulk relation between the mechanical properties of the
density, p ..... and the skeletal density, Ps [17]. The wet gel, the density and structure of the dried gel,
skeletal density was measured to be 2.15 + 0.05 and the aging parameters used during aging in the
g / c m 3 for both aged and untreated gels. The surface T E O S / E t O H solution. Increasing the aging tempera-
area decreases while the hydraulic radius increases ture causes the wet density to increase (Fig. 1) which
with increasing wet density. is caused by a higher hydrolysis and condensation
The bulk modulus of the silica network, Kp, can rate of TEOS at higher temperatures [21,22]. A
be described by a power law of the form [18,19] higher wet density is also obtained with longer aging
times caused by an increase in the amount of TEOS
= Ko( P/Po) m (2) reacted and precipitated (Fig. 1). As the strength and
The bulk modulus is related to the G modulus by stiffness of the wet gel are improved (Fig. 2), the
2(1 +
linear shrinkage during drying is reduced and a low
Kp = G 3(1 - 2 % ) ' (3) xerogel density is obtained (Fig. 1). It is of special
interest that a xerogel density of 0.24 g / c m 3 and a
where % is the Poisson's ratio of the silica network. linear shrinkage of -~ 0.5% during drying can be
The Poisson's ratio is assumed to be 0.2 [20]. Fig. obtained by increasing the wet density. The fact that
5(a) shows the bulk modulus of the wet gel versus the xerogel density reaches a minimum before it
the wet density for samples aged for different times increases again (Fig. 3) shows that the xerogel den-
and temperatures. The power law exponent of 3.4 is sity is a result of both gained weight and reduced
from a fit of all the data. Fig. 5(b) shows the bulk shrinkage during drying. As the wet density is in-
modulus of the wet gel versus the (xerogel density)/ creased, a decrease in surface area as silica precipi-
(wet density) ratio, showing that a high bulk modu- tates is observed (Fig. 4(a)). It is expected that the
lus gives xerogel densities closest to the wet density. necks between the primary particles where the silica
solubility is lowest are filled up first and second the
smallest pores [13].
4. D i s c u s s i o n In Fig. 5(a) the log-log plot demonstrates a
power-law wet density dependence of the bulk mod-
Earlier work has shown that ambient pressure ulus, Kp. Lemay et al. [23] measured the compres-
dried aerogels can be obtained by increasing the sive modulus as a function of density by preparing
mechanical strength and stiffness of the wet gel prior silica aerogels from sols with various solvent con-
S. Hcereid et aL /Journal of Non-Crystalline Solids 186 (1995) 96-103 101

centrations. The exponent was 2.9 for a base-cata- However, the precipitation of silica during aging
lyzed aerogel, 3.8 for a pre-hydrolyzed aerogel and gives an increase in strength (Fig. 1) and the increase
2.4 for an acid-catalyzed aerogel. Woignier et al. in strength is expected to be more prominent than the
[18] found a power-law behavior with an exponent of decrease in permeability [12].
3.7 of the Young's modulus for supercritical dried A simple model of structural changes in the net-
aerogels prepared under acidic, neutral, and basic work due the precipitation of silica from the aging
conditions and with various solvent concentrations. solution is developed by calculating the increase in
This work demonstrates that gels with silica precipi- cylinder radius, a, of the network and change in
tated after gelation show a similar power-law be- permeability from the cubic cell model [26]. The
haviour as gels prepared with various solvent con- cylinder radius, a, of the network can be calculated
centrations in the sol and that the exponent of 3.4 is by comparing the surface area, SA, with the cubic
in the same range as reported in the literature cell model giving [26]
[18,19,23]. As can also be seen from Figs. 5(a) and
= 1 ( 1.666- 3X)
(b), the increase in K modulus is mainly dependent
on the increase in wet density and shows no signifi- a PsSa -0.--~-3--X . (6)
cant dependence on the aging temperature. H~ereid et
If the porosity is > 30%, an approximation to give X
al. [24] have reported the increase in G modulus with
in terms of p for the cubic cell is [27]
thermal aging in water, which is dominated by the
dissolution and reprecipitation of silica, and found a 0.925
strong dependence on aging temperature. By aging in
[3/(O/Os) 'j2] - 1 (7)
a solution of monomers, the solubility of silica from
the original gel network is low and as a result aging The permeability can be estimated from [17]
in TEOS is dominated by the precipitation of silica
from the TEOS solution and not by dissolution and P rw
reprecipitation of silica from the original gel net- D= 1--- (8)
Ps 41% '
work.
Precipitation of silica during aging in the TEOS where r w is the size of the 'window' in the side of
solution gives a higher G modulus; however, it also the cell and Kw is the Kozeny constant
gives a decrease in permeability with increasing wet
Kw = 1.824 + 4.254 ( p / p ~ )
density (see Table 1). It is, however, interesting that,
even if the increase in the wet density is largest for 0.06 < ( p / p ) ~ < 0.14, (9)
between 24 and 48 h aging time, there is only a
and r w is related to the cylinder radius, a, by [17]
small decrease in permeability, indicating that the
necks between primary particles and the smallest
pores are filled up (Table 1). During further aging
(48-96 h) there is only a small difference in the wet
density, but the decrease in permeability is larger, Thus, using the calculated values of a and X, the
hence the general network thickness probably in- permeability is obtained from Eq. (8)-(10).
creases. The increase in calculated cylinder radius with
The drying stresses, orx, at the surface of a plate increasing wet density and the decrease in calculated
of gel (with thickness 2L) can be approximated by permeability are shown in Fig. 6. The calculation
[25] assumes that there is a single pore size, and all the
crx = L r k f I E / 3 D , (4) cylinders grow at the same rate. In reality, there is a
pore size distribution causing the smallest pores to
where I2e is the evaporation rate and 7k is the
fill first. However, the calculation is useful to give
viscosity of the pore liquid. In our experiment, the
an idea of the scale of the structure and indicates that
only quantity that varies is the permeability giving
the effective thickness of the polymer strands is = 1
Cx a/3-'. (5) nm. The calculated permeability (Fig. 6) is lower
102 S. Htereid et al. / J o u r n a l of Non-Crystalline Solids 186 (1995) 96-103

Table 2
Wet density, P,~.~t, xerogel density (dried from acetone), Px~o, linear shrinkage, G modulus, MOR, surface area and hydraulic radius, rh, for
samples with different treatment histories; the TEOS : EtOH ratio in the sol was 1 : 3.9 for sample (1) and 1 : 1.5 sample (2), while the water
and catalyst content in the sol were the same for both samples
Treatment history Pwct Px~,o Linear G MOR SA rh
( g / c m ~) ( g / c m -~) shrinkage (MPa) (MPa) (m2/g) (,~)
(%)
(1) Washed in H 2 0 / E t O H 24 h 60°C, 0.204 0.32 13.9 10.7 0.23 734 72
aged in T E O S / E t O H 48 h 60°C
(2) Thermal aging 48 h 60°C, 0.197 0.39 20.7 6.82 0.31 923 45
washed in H 2 0 / E t O H 24 h 60°C

than the measured permeability, probably since it is draulic radius is 72 A for xerogel (1) prepared b~¢
calculated from a single pore size. However both the aging in T E O S / E t O H solution while it is only 45 A
calculated and measured permeabilities decrease by for xerogel (2). One feasible explanation is that the
~- 50% within the same wet density range. cylinder thickness is higher for the sample aged in
A critical question concerns whether the same the TEOS solution giving a stiffer gel; the calculated
high mechanical strength and stiffness achieved by cylinder radius is 1.2 nm for sample (1), TEOS aged
precipitation of silica from the aging solution can be gel, compared with 0.94 nm for sample (2) prepared
obtained for samples with a higher silica concentra- from a sol with higher silica concentration (Eq. (6)).
tion in the sol. Gels with similar wet density, but
obtained (1) by aging an alcogel with a lower wet
density in a TEOS solution and (2) by decreasing the
5. Conclusions
ethanol content in the sol are compared in Table 2.
The gels were given the same thermal history: (1) by
aging in the TEOS solution or (2) by thermal aging (a) A gel network with a considerably higher
in the mould. Since the G modulus is higher for the mechanical strength and stiffness is obtained as silica
sample aged in TEOS, the shrinkage is reduced and precipitates from the aging solution onto the gel
hence a lower xerogel density is obtained. The hy- network causing the linear shrinkage during drying
to decrease and a lower xerogel density is obtained.
A corresponding decrease in surface area and in-
crease in hydraulic radius has been observed.
1.4 E' '' I' ''' I ' ' ' ' I ' ~' ' 3.5 (b) Shrinkage during drying can be almost com-
0 pletely eliminated. Xerogels with a density of 0.24
1.3 3.0
0 0 g / c m 3 are reported.
oo ooo oo 2.5 (c) The bulk modulus is dependent only on the
1.2 0 0o
• o--~ 2.0 amount of silica precipitated from the aging solution,
• •
1.1 not on the aging temperature.
0 0
1.5 (d) A gel prepared by precipitating silica from the
~ 1.0 aging solution shows a higher G modulus, less
1.0
shrinkage during drying, and a lower xerogel density
0.9 0.50 compared with a thermally aged gel prepared by
0.8 0.0 using a higher silica concentration in the sol.
0.10 0.15 0.20 0.25 0.30
Wet density (g/cm3) This work has been supported by the Research
Fig. 6. Calculated cylinder radius, a, and calculated permeability, Council of Norway. The authors thank Dr G.W.
D, as a function of wet density. Scherer for helpful discussions and calculations.
S. Haereid et aL /Journal of Non-Crystalline Solids 186 (1995) 96-103 103

References [13] M.-A. Einarsrud and S. H~ereid, in: Proc. 7th Int. Workshop
on Glasses and Ceramics from Gels, July 1993, J. Sol-Gel
[1] S. Haereid, Dr. ing. thesis, Norwegian Institute of Technol- Sci. Technol. 2 (1994) 903.
ogy, University of Trondheim (1993). [14] C.J. Brinker, K.D. Keefer, D.W. Schaefer, R.A. Assink, B.D.
[2] D.M. Smith, R. Desphande and C.J. Brinker, in: Better Kay and C.S. Ashley, J. Non-Cryst. Solids 63 (1984) 45.
Ceramics Through Chemistry V, ed. M.J. Hampden-Smith, [15] G.W. Scherer, in: Proc. 7th Int. Workshop on Glasses and
M.J. Klemperer and C.J. Brinker (Materials Research Soci- Ceramics from Gels, July 1993, J. Sol-Gel Sci. Technol. 2
ety, Pittsburgh, PA, 1992) p. 567. (1994) 199.
[3] G.W. Scherer and R.M. Swiatek, J. Non-Cryst. Solids 113 [16] G.W. Scherer, J. Non-Cryst. Solids 142 (1992) 18.
(1989) 119. [17] G.W. Scherer, J. Sol-Gel Sci. Technol. 1 (1994) 285.
[4] G. Scherer, J. Non-Cryst. Solids 109 (1989) 183. [18] T. Woignier, J. Phalippou and R. Vacher, J. Mater. Rcs. 4
[5] D.E. Meyers, F. Kirkbir, H. Murata, S. Ray Chaudhuri and (1989) 688.
A. Sarkar, in: Better Ceramics Through Chemistry VI, ed. C. [19] D.M. Smith, G.W. Scherer and J. Anderson, J. Non-Cryst.
Sanchez, C.J. Brinker, M.L. Mecartney and A. Cheetham Solids 188 (1995) in press.
(Materials Research Society, Pittsburgh, PA, 1994) p. 439. [20] G.W. Scherer, J. Non-Cryst. Solids 142 (1992) 18.
[6] P.J. Davis, C.J. Brinker and D.M. Smith, J. Non-Cryst. [21] M. Yamane and S. Okano, Yogyo Kyokaishi 87 (1988) 236.
Solids 142 (1992) 189. [22] M. Yamane, in: Sol-Gel Technology for Thin Films, Fibers,
[7] J.K. West, R. Nikles and G. Latorre, in: Better Ceramics Preforms, Electronics, and Speciality Shapes, ed. L.C. Klein
Through Chemistry Ill, ed. C.J. Brinker, D.E. Clark, D.R. (Noyes, Park Ridge, NJ, 1988) p. 200.
Ulrich (Elsevier, Amsterdam, 1988) p. 219. [23] J.D. Lemay, T.M. Tillotson, L.W. Hrubesh and R.W. Pekala,
[8] S.A. Pardenek, J.W. Fleming and L.C. Klein, Ultrastructure in: Better Ceramics Through Chemistry IV, ed. C.J. Brinker,
Processing of Advanced Ceramics, ed. J.D. MacKenzie and D.E. Clark, D.R. Ulrich (Materials Research Society, Pitts-
D.R. Ulrich (Wiley, New York, 1988) p. 379. burgh, PA 1990) p. 321.
[9] P.J. Davis, C.J. Brinker, D.M. Smith and R.A. Assink, J. [24] S. Ha~reid, J. Anderson, M.-A. Einarsrud, D.W. Hua and
Non-Cryst. Solids 142 (1992) 197. D.M. Smith, J. Non-Cryst. Solids 185 (1995) 221.
[10] T. Mizuno, H. Nagata and S. Manabe, J. Non-Cryst. Solids [25] C.J. Brinker and G.W. Scherer, Sol-Gel Science: The Physics
100 (1988) 236. and Chemistry of Sol-Gel Processing (Academic Press, New
[11] S. Liu and L.L. Hench, SPIE Vol. 1758 Sol-Gel Optics II York, 1990) p. 485.
(1992) 14. [26] G.W. Scherer, J. Am. Ceram. Soc. 74 (1991) 1523.
[12] S. Ha~reid, M.-A. Einarsrud and G.W. Scherer, in: J. Sol-Gel [27] G.W. Scherer, unpublished results.
Sci. Technol. 3 (1994) 199.

You might also like