Sei sulla pagina 1di 8

25908

J. Phys. Chem. B 2006, 110, 25908-25915

Engineered Complex Emulsion System: Toward Modulating the Pore Length and
Morphological Architecture of Mesoporous Silicas
He Zhang, Junming Sun, Ding Ma, Gisela Weinberg, Dang Sheng Su, and Xinhe Bao*
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian
116023, P. R. China, and Department of Inorganic Chemistry, Fritz-Haber Institute of the Max Planck Society,
Berlin D-14195, Germany
ReceiVed: September 5, 2006; In Final Form: October 18, 2006

In the complex alkane/P123/TEOS/H2O emulsion system, an emulsion engineering method to modulate pore
length and morphological architecture of mesoporous materials has been built. With fine tuning of the synthetic
parameters (e.g., the composition of the synthetic mixtures, temperature, stirring, etc.), a series of chemically
significant mesostructures (i.e., short-pore SBA-15 materials) with tunable pore length and morphological
architecture have been successfully constructed. The effects of alkane solubilizates on pore length and particle
morphology are discussed. The resulting short-pore materials would have potential applications in the fields
of adsorption/separation of biomolecules and inclusion chemistry of guest species, etc.

1. Introduction
During the past few decades, supermolecular self-assembly
has initiated a new era, in which the controlled bottom-up
construction of various functional porous materials has become
the interest of many researchers. As a promising support, the
construction of ordered mesoporous silicas is always one of the
research focuses in this newly developed field.1 Almost along
with the discovery of M41s,2 studies to control the morphology/
texture of these obtained mesoporous materials have begun,
aiming at getting good performance in the specific applications.
Due to the existence of complicated interactions among the
different species in the synthetic mixtures, a slight change in
synthetic parameters, such as temperature, shearing, components,
additives, etc., might alter the self-assembly behavior significantly. Not surprisingly, mesoporous materials with diverse
morphologies/textures resulted.3-13
The construction of chemically significant mesostructures has
long been one of the hottest issues in the supermolecular selfassembly age.6 Particularly, mesoporous materials with short
and open channels are extensively investigated because they
are more favorable for the mass transfer. For example, in the
adsorption experiments of biomolecules, due to the long, curved
mesostructures and densely packed secondary morphologies of
traditional SBA-15,14 the diffusion of large biomolecules (e.g.,
lysozyme) within their channels is suppressed, and thus, the
adsorption equilibrium was often reached on a time scale of
more than 24 h. If SBA-15 with straight pore channels and
separated rodlike morphologies was used, significant improvement in the diffusion of lysozyme molecules was observed, and
adsorption equilibrium time was shortened to a few hours.15
More recently, by using nanoscale large-pore mesoporous silicas,
the time scale for lysozyme adsorption equilibrium has been
further decreased to a few minutes.16 It was also demonstrated
that more accessible mesopores will result in a significant
increase in the amount of lysozyme adsorption. All these results
* Towhomcorrespondenceshouldbeaddressed.E-mail: xhbao@dicp.ac.cn.
Fax: 86-411-84694447.
Chinese Academy of Sciences.
Fritz-Haber Institute of the Max Planck Society.

indicate that the design and synthesis of chemically significant


morphological architectures are very important for the practical
applications of mesoporous materials.
To get short and open porous structures, decreasing the size
of the particles was the most used method, by changing
quenching procedure, dilution of the reaction solutions, spray
drying, as well as a secondary surfactant mediated process,
etc.17-20 These methods were mainly focused on the synthesis
of the cationic surfactants directed MCM-41 nanoparticles. On
the other hand, controlling the length of surfactant micelles21-23
could also result in the short-pore materials; however, few
reports were successful until now.24 Since 1998, the nonionic
surfactant-templated mesoporous SBA-15 silicas have attracted
more and more attention because of their larger surface areas,
tunable pore size, thicker pore walls, and therefore higher
hydrothermal stability.14 However, as mentioned above, the long
end to end and side by side densely packed morphologies of
the traditional SBA-15 silicas are not suitable for many practical
applications, especially where a fast mass transfer is required.
Although various strategies have been developed for the
modulation of their morphological architectures, limited cases
about the construction of chemically significant structures20 (e.g.,
nanosized particles) were reported. Even for these successful
cases, the sacrifice of ordering degree of mesostructures is often
observed. In our previous studies, large amounts of decane were
used to modulate the length of P123 micelles; as a result, unusual
SBA-15 materials with parallel channels running along the short
axis have been constructed.24 Our recent studies show that the
interaction between different alkanes and P123 micelles are not
the same.25 By strictly controlling the initial reaction temperature, alkanes with shorter chain length (e.g., hexane) could be
used to construct highly ordered SBA-15.25 On the basis of this
knowledge, herein, by finely tuning the synthetic parameters
(e.g., the composition of the synthetic mixtures, temperature,
stirring, etc.) of the alkane (from pentane to hexadecane)/P123/
water/TEOS complex emulsion systems, a series of well-ordered
chemically significant mesostructures with tunable pore length
and morphological architectures, from uniform submicrometersized SBA-15 columns to short-pore SBA-15 hexagonal slices

10.1021/jp065760w CCC: $33.50 2006 American Chemical Society


Published on Web 11/30/2006

Mesoporous Silicas

J. Phys. Chem. B, Vol. 110, No. 51, 2006 25909

TABLE 1: Parameters of SBA-15 Materials Prepared by Using Different Alkanes


alkane

initial reaction temp (K)

D100 spacing (nm)

unit cell (a0, nm)

pore diameter (nm)

surface area (m2/g)

none
hexadecane
dodecane
decane
nonane
octane
heptane
hexane
pentane

315 ( 2
316 ( 2
313 ( 2
308 ( 2
300 ( 2
298 ( 2
293 ( 2
288 ( 2
285 ( 2

10.1
11.6
11.6
11.6
12.3
13.0
13.4
14.2
14.6

11.7
13.4
13.4
13.4
14.2
15.0
15.5
16.4
16.8

9.7
12.0
12.0
12.0
13.4

659
539
600
560
637

15.0
15.7

620
614

to mesoporous free-standing films and monoliths with shortpore SBA-15 units, have been successfully constructed. At the
same time, the effects of alkanes on pore length and particle
morphology are discussed, and possible mechanisms for the
formation of these short-pore materials are proposed. Significantly, most of these short-pore mesoporous silicas have shown
superiority against traditional SBA-15 in the fields of biomolecular adsorption, separation, and inclusion of guest species.16,26
2. Experimental Methods
2.1. Synthesis. As a typical synthesis procedure, 2.4 g of
EO20PO70EO20 (P123) was dissolved in 84 mL HCl solution
(1.20 M), followed by the addition of 0.027 g of NH4F. The
mixture was then stirred at a given temperature (from 285 to
315 K) until the solution became clear. Different alkanes (from
pentane to hexadecane) and TEOS were premixed and then
introduced into the solution under mechanical stirring (300360 rpm) (final P123/HCl/NH4F/H2O/TEOS/alkane molar ratios
) 1/261/1.8/11278/x/y; x ) 48-110, y ) 0-755). The above
mixture was stirred at the given temperature (Table 1) for 20 h,
and then transferred into an autoclave for further reaction at
373 K for 48 h. The products were collected by filtration, dried
in air, and calcined at 813 K for 5 h to remove the templates.
2.2. Characterization. SEM was done on the Hitachi S4800
field-emisson scanning electron microscope. The TEM images
were obtained with a Philips CM 200 transmission electron
microscope equipped with a CCD camera.
XRD patterns were collected on a Rigaku D/MAX 2400
diffractometer equipped with a Cu KR X-ray source operating
at 40 kV and 50 mA. The N2 adsorption-desorption isotherms
were recorded on an ASAP 2000 instrument.
3. Results
3.1. Effect of Alkane Solubilizates on the Pore Length of
SBA-15. Various alkanes from pentane to hexadecane (final
P123/HCl/NH4F/H2O/TEOS/alkane molar ratio ) 1/261/1.8/
11278/60/235) have been used in the synthesis. Figure 1 shows
SAXD patterns of the samples prepared with different alkanes.
It is clear that all the samples show four well-resolved peaks
that can be indexed as (100), (110), (200), (210) diffractions
associated with a 2-D hexagonal symmetry (p6mm), characteristic of highly ordered SBA-15 silicas. In addition, shift of the
(100) diffraction to a low angle with the decrease of alkane
chain length indicates the increase of SBA-15 unit cell.25 More
physical parameters of the obtained SBA-15 materials can be
found in Table 1.
Low magnified SEM images (Figure S1) show that the
morphological architectures of the obtained SBA-15 are different
depending on the alkane solubilizates used. Most importantly,
HRSEM images at higher magnification demonstrate that the
pore length of the obtained SBA-15 could be modulated by using
different alkanes (Figure 2). Without alkane, the obtained
materials have a bundle morphology, which consists of long

ropelike particles (above 1 m) with parallel channels running


along the long morphological axis (Figure 2a). When dodecane
is used, the long ropelike morphology is maintained, except that
the particles turn thinner (ca. 1 m, Figure 2b). When decane
is used, however, the morphology changes into uniform columnlike particles, with a pore length of ca. 260 ( 50 nm (Figure
2c). When the alkane chain length is further decreased (n-C9,
Figure S1e), the morphology of the obtained SBA-15 changes
back into long rodlike particles again (ca. 700-900 nm long).
Surprisingly, the pores of the obtained SBA-15 are running along
the short morphological axis with a pore length of only ca. 150
( 50 nm (Figure 2d). Upon decreasing the alkane chain length
to n-C8 (octane), the structure with parallel channels running
along the short axis is maintained, but the pore length increases
to ca. 210 ( 50 nm (Figure 2e). When heptane is used, the
morphology of the obtained materials changes significantly, and
micrometer-scale single or twin bunches were obtained. Moreover, the TEM image (inset in Figure S1g) shows that the
micrometer-sized bunches are in fact composed of fused shortpore SBA-15 particles (<300 nm). When hexane is used,
ropelike SBA-15 particles, which are similar to those conventional SBA-15 silicas, are again obtained, and the parallel
channels are along the long morphological axis (Figure 2f).
Figure 3 summarizes the pore length as a function of alkane
chain length. It is clear that, with the decrease of alkane chain
length, the pore length of the obtained SBA-15 decreases and
goes through a minimum, and then increases again. At the same
time, it is observed that the pore size increases with the decrease
of alkane chain length.25
In the complex alkane/P123/TEOS/H2O emulsion system, for
the first time, it is demonstrated that the introduction of alkane
solubilizates could tune the pore length of SBA-15. At the same
time, the particle morphologies of the obtained materials are
different depending on the alkane solubilizates used. Significantly, using alkanes with medium chain length (i.e., from

Figure 1. Small-angle X-ray diffraction patterns of the obtained SBA15 synthesized by using (a) hexane, (b) octane, (c) nonane, (d) dodecane,
(e) no alkane.

25910 J. Phys. Chem. B, Vol. 110, No. 51, 2006

Zhang et al.

Figure 2. HRSEM images of the obtained SBA-15 mesoporous silicas resulting from the use of different alkanes as solubilizates: (a) no alkane,
(b) dodecane, (c) decane, (d) nonane, (e) octane, (f) hexane.

Figure 3. Effects of carbon number of alkanes used on pore length of


obtained mesoporous materials.

heptane to decane) in the synthesis has led to a series of shortpore SBA-15 with chemically significant morphologies.
3.2. Effect of the Amount of Decane on Particle Morphology. Keeping the TEOS/P123 molar ratio at 60, the effect of
the amount of decane on the particle morphology is also
investigated (final P123/HCl/NH4F/H2O/TEOS/decane molar
ratios ) 1/261/1.8/11278/60/x, x ) 0-755). Small-angle X-ray
diffraction patterns (Figure S2) of mesoporous silicas prepared
with different decane/P123 molar ratios show four well-resolved

peaks that can be indexed as (100), (110), (200), and (210)


diffractions associated with a 2-D hexagonal symmetry (p6mm).
Bundles of silica fibers are the main products (Figure S1a) in
the absence of decane.27 When the decane/P123 molar ratio
equals 134, long fiberlike materials (micrometer size) account
for more than 95% of the products (Figure S3). The SEM image
(Figure 4a) shows that the long fibers are actually composed
of particles bunches, the diameter of which is ca. 300 nm, much
smaller than that of the fiberlike SBA-15 particles prepared
without decane. It suggests that decane could indeed tune the
length of P123 micelles during the synthesis.24 The TEM image
(Figure S3) shows that the ordered mesoporous particles are
imperfectly attached to each other along the [001] direction,
forming the SBA-15 particles bunches, which are similar to the
reported towerlike SBA-15.28
When the amount of decane is increased (decane/P123 )
235), the SBA-15 bunches completely disappear, and isolated
particles with a yield of more than 90% are observed (Figure
4b). The morphology of a single particle is similar to the
building block of SBA-15 bunches prepared with the decane/
P123 molar ratio of 134. Further increasing the amount of
decane (decane/P123 ) 755), lots of monoliths are observed.
Interestingly, these monoliths are composed of submicrometersized short-pore particles, which may be formed at continuous
oil/water interfaces (Figure 4c,d).

Mesoporous Silicas

J. Phys. Chem. B, Vol. 110, No. 51, 2006 25911

Figure 4. SEM images of mesoporous silicas prepared with a 60 TEOS/P123 molar ratio and different decane/P123 molar ratio: (a) 134, (b) 235,
(c, d) 755.

Figure 5. Small-angle X-ray diffraction patterns of SBA-15 prepared


with a 235 decane/P123 molar ratio and different TEOS/P123 molar
ratio: (a) 48, (b) 60, (c) 77, (d) 110.

3.3. Effect of the Amount of TEOS on Particle Morphology. Fixing the decane/P123 molar ratio at 235, the effect of
the TEOS/P123 molar ratio on particle morphology is investigated (final P123/HCl/NH4F/H2O/TEOS/alkane molar ratios )
1/261/1.8/11278/y/235; y ) 48-110) in the current study. Figure
5 shows the small-angle X-ray diffraction patterns of the
samples. All the samples show diffraction peaks associated with
a 2-D hexagonal symmetry (p6mm), which is consistent with
the N2 adsorption results (Figure S4). For the sample with the
TEOS/ P123 molar ratio of 48, although it gives only two broad
peaks, it still possesses a well-ordered mesostructure, and fits a
minimum hexagonal mesostructure arrangement of ultrafine
particles, which could be verified by TEM experiments (Figure
6b).16 Corresponding SEM demonstrates a fiberlike bundle
morphology (Figure 6a). The TEM image in Figure 6b shows
the products are well-aligned nanofibers (50-75 nm), and each
of the fibers is indeed in a hexagonal arrangement. When the
TEOS/P123 molar ratio equals 60 (Scheme 1), uniform meso-

porous columns are obtained (Figure 1c). Furthermore, the crosssection TEM image (Figure S5) shows that the obtained
materials have hexagonal arrangements, which is consistent with
the SAXD results. Further increasing the molar ratio of TEOS
to P123 (e.g., 77, Scheme 1), interestingly, large amounts of
well-defined hexagonal-shaped SBA-15 slices are observed,
which are similar to the slice-shaped organosilicas29 (Figure 6c,
more pictures could be found in Figure S7). HRSEM images
(Figure S6) show that the ordered parallel pore channels are
vertical to the (100) planes of the slices, and the length of the
pore channels is about 200 nm, which is shorter than that of
the SBA-15 columns prepared with a TEOS/P123 molar ratio
of 60, while the dimension along the (100) plane has increased
to ca. 1 m. Microtome technique has also been used to confirm
the nanostructures of SBA-15 slices. Figure 6d gives the crosssection image vertical to the (100) planes. It can be clearly seen
that the ordered parallel channels are running along the [001]
direction. Another TEM image in Figure S6 shows the
hexagonal pore arrays of SBA-15 slices taken along the pore
direction ([001] direction). At the same time, it is observed that
the aspect ratio (the ratio of the particle dimension along the
parallel channels to that vertical to the parallel channels) of
obtained particles decreases with the increase of the TEOS/
P123 molar ratio.
Further increasing the amount of TEOS (e.g., TEOS/P123 )
110), large amounts of free-standing films (more than 50% in
yields) are observed in the obtained products (Figure 6e,f). The
formation of the films could be attributed to the oil droplets
directed macroscale growth in the synthetic mixtures. However,
HRSEM images (more images can be found in Figure S7) show
that the orientation of the pore channels is running parallel to
the surface of the films, which is different from that obtained
with CTAB surfactant.30 This difference is possibly related to
the different interactions between oil and surfactants. Besides

25912 J. Phys. Chem. B, Vol. 110, No. 51, 2006

Zhang et al.

Figure 6. SEM images (a, c, e, f) and TEM images (b, d) of mesoporous silicas prepared with a 235 decane/P123 molar ratio but different
TEOS/P123 molar ratios: (a, b) 48, (c, d) 77, (e, f) 110.

SCHEME 1: Schematic Formation of Different


Morphological SBA-15 Particles

latter is assembled with large amounts of short-pore subunits.


Both mesostructures are believed to be formed on the oil/water
interfaces though with different morphologies. The formation
mechanism of different architectures is still unclear. We propose
that it could be related to the difference of TEOS concentration
and silicate oligmers.
4. Discussion

the mesoporous films, some aggregated irregular particles are


also observed (Figure S8).
It should be mentioned that, herein, the architecture of the
obtained mesoporous films is different from that of the monoliths
formed in the presence of large amounts of alkanes (Figure 4d).
The former is a smooth and continuous integration, while the

4.1. Pore Length Evolution of SBA-15 with the Alkane


Chain Length. It is well-known that the equilibrium properties
of the surfactant aggregates (e.g., the shape and size of the
micelle) are determined by the variation of the standard chemical
potential of the surfactant with different shapes and aggregation
numbers.31 The effect of solubilizates on the shape and size of
ionic micelles has been investigated extensively.21-23 Generally,
it is widely accepted that the effect of an additive (solubilizate)
on the aggregation could be correlated to its solubilization sites.
One typical example is the solubilization of aromatic molecules
in the cationic surfactants micelles. The strong interactions
between the electrons of aromatics and positively charged
headgroups of the surfactants make the aromatic molecules
dissolved preferentially at or near the interfaces of the aggregates, lowering the electrostatic repulsion between the
headgroups and the hydrocarbon/water interfacial free energy.
Because of this, a proper amount of aromatic molecules often

Mesoporous Silicas

J. Phys. Chem. B, Vol. 110, No. 51, 2006 25913

Figure 7. TEM images of mesoporous silicas prepared at 313 K with different alkanes: (a) dodecane, (b) heptane, (c) hexane.

result in the formation of ultra-long micelles. A further increase


in the amount of aromatic solubilizates, however, saturates the
outer layer of micelles, which is followed by a further
solubilization within the cores of micelles with a consequent
axial ratio shortening of rod micelles and further transformation
from rodlike to globular aggregates.22-23 The solubilization of
aliphatic hydrocarbons in the traditional ionic surfactant micelles
(e.g., C16TABr) shows two different trends in tuning the size
and shape of aggregates depending on their chain length.32
However, solubilization of the alkane-induced rod-to-sphere
transition was thought to be the consequence of the axial ratio
shortening of rod micelles.21
In the block copolymer (e.g., P123)/water binary systems,
solubilization of alkanes is mainly influenced by the interactions
betweenthesolubilizatesandhydrophobicblocksofcopolymers,33-35
which is obviously different from that in the ionic surfactant
micelles where interfacial energy determines the nature of
solubilization. According to Nagarajan et al.,33 the alkanes would
be predominantly solubilized in the hydrophobic cores of the
P123 micelles, which add a free energy contribution, shift the
delicate balance of intermolecular forces, and thus determine
equilibrium properties of micelles (e.g., shape and size of
micelles). Moreover, the interactions between alkanes and PPO
blocks are different depending on the chain length of alkanes.33
At a given synthesis temperature, the shorter the chain length
of the alkane is, the stronger the interaction between alkanes
and PPO blocks would be, and more alkanes are able to be
accommodated in the cores of micelles.25 For example, when
the decane/P123 molar ratio is 235, the pore length of SBA-15
becomes shorter and the corresponding structure transforms to
MCFs (globular shape) with the decrease of alkane chain length
at a given reaction temperature (e.g., 313 K) (Figure 7). This
result indicates that the solubilization ability of alkanes indeed
affects the equilibrium properties of micelles greatly. At a fixed
temperature, with the decrease of alkane chain length, more
alkane would be solubilized in the PPO block domain inside
the micellar cores. It leads to the axial ratio shortening of rodlike
micelles (herein, we named it as positive effect) and finally the
transition from rodlike to spherical micelles, consistent with the
literature results.21 On the other hand, however, our previous
studies showed that for shorter-chain alkanes (e.g., hexane), only
by decreasing the initial reaction temperatures, can we get highly
ordered SBA-15,25 which suggests that the initial reaction
temperature is also important for the control of the equilibrium
properties of micelles. Moreover, decreasing reaction temperature would result in the recovery of rodlike micelles from those
spherical ones. It indicates that lower reaction temperatures favor
the elongation of rodlike micelles (herein, we named it as
negative effect).
Herein, both the solubilization of alkanes and the initial
reaction temperatures govern the pore length of obtained SBA15 synthesized with the alkanes of different chain lengths. When

an alkane with a longer chain (e.g., dodecane) was used, a small


amount of alkane could be solubilized in the micellar cores of
P123 surfactant, and no obvious effect could be observed on
the pore length of SBA-15 (Figure 3). With the decrease of
alkane chain length, the solubilization ability of alkane increases.25,33 As a result, the pore length of SBA-15 decreases
accordingly. By further decreasing the alkane chain length
(shorter than n-C9), in order to get highly ordered mesoporous
SBA-15, however, the initial reaction temperature employed is
also decreased significantly.25 At this stage, we propose that
the reaction temperature would begin to play the predominant
role in controlling the pore length. The pore length of SBA-15
thus increased again with the decrease of alkane chain length.
It must be kept in mind that the temperature also has a great
influence on the solubilization ability of alkanes in the block
copolymers micelles. The lower the temperature is, the smaller
amount of alkane can be solubilized in the block copolymers
micelles.36 Therefore, temperature has some effects on the
solubilization ability of alkanes and thus the pore length of SBA15. However, our previous results indicate that the solubilization
ability increases continuously with the decrease of alkane chain
length even at a lower reaction temperature.25 Therefore, in this
case, the effect of reaction temperature on the solubilized ability
of alkanes and thus the pore length of SBA-15 could be
excluded. The detailed mechanism of the negative effect by
reaction temperature is still unclear. We propose that it must
be related to the special temperature-sensitive character of
copolymers. It should also be mentioned that the existence of
large amounts of alkanes plays a key role in tuning the pore
length of SBA-15. As mentioned above, the rod-to-sphere
transition of micelles is the consequence of the decreasing pore
length of micelles. A small amount of alkanes would be not be
sufficient to cause such a transition and thus not the axial ratio
shortening of micelles.24 However, it still has an obvious effect
on the morphology of the obtained SBA-15 silicas.
4.2. Formation of Bunches, Columns, and Monoliths of
Short-Pore SBA-15. Triblock copolymers, such as pluronic
P123 surfactant with the proper PO/EO ratios, have been shown
to aggregate in the form of long micelles. Together with a proper
shearing force, ropelike or fiberlike morphological SBA-15
aligned side-by-side and end-to-end often result. Under static
conditions, however, micrometer-sized SBA-15 with a rodlike
morphology would result.36 It is also reported that the different
surface energy between the hydrophobic core (PPO) and
hydrophilic corona (PEO) of the silicate-doped micelles in the
synthetic aqueous solutions was proposed to be the driving force
in the oriented attachment of the SBA-15 particles at the flat
basal (100) planes into towerlike SBA-15.28 In this case, the
difference of surface energy must be taken into account,
particularly, when large amounts of decane are introduced in
the current system. Our studies indicated that when a relatively
small amount of decane (decane/P123 < 70) was used, it would

25914 J. Phys. Chem. B, Vol. 110, No. 51, 2006

Zhang et al.

Figure 8. SEM images of SBA-15 materials prepared in the absence of alkane solubilizates at a TEOS/P123 molar ratio of (a) 48, (b) 60, (c) 77.

attach preferably on the relatively hydrophobic cores of the


micelles, resulting in the expansion of pore size of SBA-15.24
Further increasing the amount of decane (decane/P123 ) 134)
does not further increase the pore size of SBA-15. Therefore, it
can be concluded that the amount of decane (decane/P123 )
235) used in the current study is more than enough to expand
the size of the pores, and large amounts of oil droplets would
be present in the synthetic mixtures, which would act as spacer
in the current system. We propose that the compartmentalized
effect of the emulsion droplets or the preferential interaction
between the oil droplets and the hydrophobic basal (100) planes
of the SBA-15 particulates suppressed the oriented attachment
of the SBA-15 particles to each other along the [001] direction.
When decane/P123 equals 134, imperfect attached SBA-15
bunches are obtained (Figure 4a), suggesting that that amount
of alkane is not sufficient to suppress the aggregation of SBA15 particles completely (Scheme 1). Only when excess amounts
of decane were used (e.g., decane/P123 ) 235) could the
oriented arrangement of SBA-15 particles along the [001]
direction be destroyed completely (thus, isolated SBA-15
columns are obtained (Scheme 1)). With an increase in the
amount of decane (e.g., decane/P123 ) 755), the monoliths
assembled with such short-pore columns result at the oil-water
interfaces, which result from the existence of large amounts of
decane.
4.3. Formation of Short-Pore SBA-15 Particles with the
Tunable Aspect Ratios. In contrast to the alkane solubilizates
which are mainly located within the micellar cores (subsection
4.1), silicate species would preferentially stay at the outer surface
of the micelles. Moreover, the amount of TEOS determines the
type and concentration of silicate species in the systems, which
could affect their binding strength with P123 micelles and
aggregation capabilities of inorganic-organic composites.37
Significantly, the charge density at the polar part of the micelles
would also be affected,38-39 which would have an important
effect on the length of the rodlike micelles.22-23 In the current
study, it could be proposed that, with the increase of the TEOS/
P123 molar ratio, larger silicate oligomers would be created
and attached at the outer EO brushes of the P123 micelles during

the synthesis. As a result, the aggregation capabilities of rodlike


micelles increase, and the side-on condensations of the obtained
materials are thus preferred. On the other hand, the charge
density at the outer surface of P123 micelles increases too, and
the electrostatic repulsion between the charged headgroups
increases accordingly,38 which undoubtedly leads to the decrease
of the rod micelle length. Indeed, keeping the decane/P123 molar
ratio at 134, and increasing the TEOS/P123 molar ratio from
48 to 77, the aspect ratio of the obtained materials decreases
from 4.5 to 0.18, further confirming the validity of our proposal.
Further increasing the decane/P123 molar ratio to 110, large
amounts of free-standing mesoporous films result, the reason
for which is unclear. But the formation of these free-standing
mesoporous films should be involved in a self-assembly process
at the interfaces of the decane (or the mixture of decane and
TEOS) and water.
It must be kept in mind that the decane also plays a very
important role in controlling the length of the micelles and the
aspect ratio of the obtained SBA-15 during the experiments of
tuning the TEOS amounts. To understand the function of decane,
the same TEOS/P123 molar ratio (from 48 to 77) has been used
to prepare different samples in the absence of decane. From
SAXD patterns of the samples (Figure S9), they have highly
ordered hexagonal mesostructures. SEM images (Figure 8) show
that although the aspect ratio of the obtained SBA-15 decreased
with the increasing amount of TEOS, however, compared with
that prepared in the presence of decane, the decrease of the
aspect ratio is much smaller, i.e., from 7.4 to 1.7, suggesting
that the aspect ratio of SBA-15 materials is governed by both
the alkane solubilizates and the amount of TEOS (Figure 9).
Interestingly, similar results have been recently reported by
increasing the Si/P123 ratio in the P123/organic silica source
(BTMSE)/HCl system,29 indicating that organic functional
groups could play a positive role in decreasing the aspect ratio
of particles.
5. Conclusions
In the complex alkane/P123/TEOS/H2O emulsion system, an
emulsion engineering methodology to modulate the pore length

Mesoporous Silicas

Figure 9. Effects of TEOS/P123 molar ratio with a 235 decane/P123


molar ratio on the aspect ratio of particles.

and morphological architectures of mesostructured materials has


been built. By controlling the alkane solubilizates, the amount
of TEOS added, and other parameters (e.g., temperature), it is
possible to tune the pore length and morphology of the obtained
SBA-15 silicas. As a result, SBA-15 with chemically significant
morphologies, such as columns, slices, cuboids, nanoscale rods,
as well as monoliths assembled with short-pore SBA-15 silica
particles, etc., were constructed. Most importantly, some of these
materials (e.g., SBA-15 columns) exhibit high efficiency and
selectivity as the stationary phases of capillary electrochromatography (CEC),26 and may have many potential applications
in other fields, such as the separation of biomolecules and the
inclusion of guest species.
Acknowledgment. We are grateful for the support of
National Natural Science Foundation of China (No. 90206036)
and the Ministry of Science and Technology of China
(2005CB221405).
Supporting Information Available: Low magnified SEM
images of SBA-15 materials prepared in the presence of large
amounts of different alkane solubilizates. (HR)SEM/TEM
images of SBA-15 bunches, SBA-15 columns, hexagonal SBA15 slices, and free-standing SBA-15 films. Small-angle X-ray
diffraction patterns of SBA-15 materials prepared with different
decane/P123 and TEOS/P123 molar ratios without decane. N2
adsorption and desorption isotherms of SBA-15 columns and
hexagonal SBA-15 slices. This material is available free of
charge via the Internet at http://pubs.acs.org.
References and Notes
(1) Taguchi, A.; Schuth, F. Microporous Mesoporous Mater. 2005, 77,
1.
(2) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C. Beck,
J. S. Nature 1992, 359, 710.

J. Phys. Chem. B, Vol. 110, No. 51, 2006 25915


(3) Chen, B. C.; Lin, H. P.; Chao, M. C.; Mou, C. Y.; Tang, C. Y.
AdV. Mater. 2004, 16, 1657.
(4) Kim, T. W.; Kleitz, F.; Paul, B.; Ryoo, R. J. Am. Chem. Soc. 2005,
127, 7601.
(5) Sun, J. M.; Ma, D.; Zhang, H.; Wang, C.; Bao, X. H.; Su, D. S.;
Klein-Hoffmann, A.; Weinberg, G.; Mann, S. J. Mater. Chem. 2006, 16,
1507.
(6) Karkamkar, A. J.; Kim, S.; Mahanti, D.; Pinnavaia, T. AdV. Funct.
Mater. 2004, 14, 507.
(7) Wang, J.; Zhang, J.; Asoo, B. Y.; Stucky, G. D. J. Am. Chem. Soc.
2003, 125, 13966.
(8) Yao, B. D.; Fleming, D.; Morris, M. A.; Lawrence, S. E. Chem.
Mater. 2004, 16, 4851.
(9) Zhao, D. Y.; Sun, J. Y.; Li, Q. Z.; Stucky, G. D. Chem. Mater.
2000, 12, 275.
(10) Huo, Q.; Feng, J.; Schuth, F.; Stucky, G. D. Chem. Mater. 1997,
9, 14.
(11) Yu, C. Z.; Fan, J.; Tian, B. Z.; Zhao, D. Y. Chem. Mater. 2004,
16, 889.
(12) Sayari, A.; Han, B. H.; Yang, Y. J. Am. Chem. Soc. 2004, 126,
14348.
(13) Lin, H.; Mou, C. Acc. Chem. Res. 2002, 35, 927.
(14) Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G. H.;
Chmelka, B. F.; Stucky, G. D. Science 1998, 279, 548.
(15) Fan, J.; Lei, J.; Wang, L. M.; Yu, C. Z.; Tu, B.; Zhao, D. Y. Chem.
Commun. 2002, 2140.
(16) Sun, J. M.; Zhang, H.; Tian, R. J.; Ma, D.; Bao, X. H.; Su, D. S.;
Zou, H. F. Chem. Commun. 2006, 1322.
(17) Cai, Q.; Luo, Z.; Pang, W.; Fan, Y.; Chen, X.; Cui, F. Chem. Mater.
2001, 13, 258.
(18) Nooney, R. I.; Thirunavukkarasu, D.; Chen, Y.; Josephs, R.; Ostafin,
A. E. Chem. Mater. 2002, 14, 4721.
(19) Suzuki, K.; Ikari, K.; Imai, H. J. Am. Chem. Soc. 2004, 126, 462.
(20) Han, Y.; Ying, J. Y. Angew. Chem., Int. Ed. 2005, 44, 288.
(21) Heindl, A.; Strnad, J.; Kohler, H. H. J. Phys. Chem. 1993, 97, 742.
(22) Tornblom, M.; Henriksson, U. J. Phys. Chem. B 1997, 101, 6028.
(23) Hoffmann, H.; Ulbricht, W. J. Colloid Interface Sci. 1989, 129,
388.
(24) Zhang, H.; Sun, J. M.; Ma, D.; Bao, X. H.; Klein-Hoffmann, A.;
Weinberg, G.; Su, D. S.; Schlogl, R. J. Am. Chem. Soc. 2004, 126, 7440.
(25) Sun, J. M.; Zhang, H.; Ma, D.; Chen, Y.; Bao, X. H.; KleinHoffmann, A.; Pfander, N.; Su, D. S. Chem. Commun. 2005, 5343.
(26) Tian, R. J.; Sun, J. M.; Zhang, H.; Ye, M. L.; Xie, C. H.; Dong, J.;
Hu, J. W.; Ma, D.; Bao, X. H.; Zou, H. F. Electrophoresis 2006, 27, 742.
(27) Schmidt-Winkel, P.; Yang, P. D.; Margolese, D. I.; Chmelka, B.
F.; Stucky, G. D. AdV. Mater. 1999, 11, 303.
(28) Fahn, Y. Y.; Su, A. C.; Shen, P. Langmuir 2005, 21, 431.
(29) Bao, X. Y.; Zhao, X. S. J. Phys. Chem. B 2005, 109, 10727.
(30) Schacht, S.; Huo, Q.; Voigt-Martin, I. G.; Stucky, G. D.; Schuth,
F. Science 1996, 273, 768.
(31) Nagarajan, R.; Rukenstein, E. Langmuir 1991, 7, 2934.
(32) Lianos, P.; Lang, J.; Sturm, J.; Zana, R. J. Phys. Chem. 1984, 88,
819
(33) Nagarajan, R.; Barry, M.; Ruckenstein, E. Langmuir 1986, 2, 210.
(34) Nagarajan, R. Curr. Opin. Colloid Interface Sci. 1997, 2, 282.
(35) Forster, S.; Antonietti, M. AdV. Mater. 1998, 10, 195.
(36) Kosuge, K.; Sato, T.; Kikukawa, N.; Takemori, M. Chem. Mater.
2004, 16, 899.
(37) Chao, M.; Lin, H. P.; Sheu, H.; Mou, C. Y. Stud. Surf. Sci. Catal.
2002, 141, 387.
(38) Huo, Q.; Margolese, D. I.; Ciesla, U.; Demuth, D.; Feng, P.; Gier,
T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schuth, F.; Stucky, G. D.
Chem. Mater. 1994, 6, 1176.
(39) Antonietti, M.; Goltner, C. Angew. Chem., Int. Ed. 1997, 36, 910.

Potrebbero piacerti anche