Sie sind auf Seite 1von 6

ARTICLE

pubs.acs.org/Langmuir

Rapid Synthesis of SBA-15 Rods with Variable Lengths, Widths, and


Tunable Large Pores
Emma M. Johansson,* Mohamed A. Ballem, Jose M. Cordoba, and Magnus Oden
Nanostructured Materials, Department of Physics, Chemistry and Biology, Linkoping University, Linkoping SE-58183, Sweden
ABSTRACT: Dispersed SBA-15 rods have been synthesized
with varying lengths, widths, and pore sizes in a low-temperature synthesis in the presence of heptane and NH4F. The pore
size of the material can systematically be varied between 11 and
17 nm using dierent hydrothermal treatment times and/or
temperatures. The particle length (400600 nm) and width
(100400 nm) were tuned by varying the HCl concentration.
All the synthesized materials possess a large surface area of
400600 m2/g and a pore volume of 1.051.30 cm3. A
mechanism for the eect of the HCl concentration on the particle morphology is suggested. Furthermore, it is shown that the
reaction time can be decreased to 1 h, with well-retained pore size and morphology. This work has resulted in SBA-15 rods with the
largest pore size reported for this morphology.

INTRODUCTION
Ordered mesoporous silica can be synthesized in a variety of
pore sizes, pore shapes, pore arrangements, and morphologies.
Due to its large surface area, a material with a controllable pore
size and a narrow pore size distribution has many potential
applications, such as in separation, catalysis, and adsorption and
as templates.14 This class of materials has received much
interest in the past few decades, and depending on the application, various morphologies, pore sizes, and pore shapes are
preferable.
SBA-155,6 with its hexagonally ordered cylindrical pores can
be synthesized in a variety of morphologies, e.g., bers,7 spheres,8
platelets,9 or rods.10 The rodlike morphology of SBA-15 is of
interest due to its small dimensions, which result in short
diusion paths and thereby the possibility of fast adsorption
and mass transfer. This has been proven useful in applications
such as immobilization of enzymes1113 or as a template for
mesoporous carbon.14 The key to synthesizing rods is shortening
the stirring time from the normal 220 h to less than 10 min.
Rods with 69 nm large pores have been synthesized without
additives,10 in the presence of KCl14 or glycerol,15 or with sodium
metasilicate as the silica precursor.16 The length of these rods is
normally 12 m, but by varying the HCl concentration, the
length can be tuned from 0.3 to 4 m when glycerol is present
during the reaction.15
The pore size of SBA-15 is normally 69 nm, but it can be
increased by, e.g., adding swelling agents such as 1,3,5-trimethylbenzene (TMB)6 or by varying the hydrothermal treatment time
and temperature.17 It is possible to increase the pore size to
12 nm using swelling agents. Further increase was limited by a
phase transition from ordered hexagonal pores to disordered
mesocellular foams.18 Lately it has been shown that low-temperature syntheses with alkanes and NH4F can increase the pore
r 2011 American Chemical Society

size above 12 nm,19,20 where the morphology can be altered by


varying synthesis parameters such as tetraethyl orthosilicate
(TEOS)/triblock copolymer EO20PO70EO20 (P123) or alkane/P123 molar ratios. We have recently shown that by varying
the heptane to P123 molar ratio it is possible to synthesize SBA15 with varying morphologies and with controlled pore sizes of
up to 18 nm.21,22
In the search to have more eective synthesis conditions, the
time between stirring and the hydrothermal treatment was
studied. It has previously been shown that regular syntheses of
SBA-15 can have a reduced synthesis stirring time from 20 to 2 h
without compromising the quality of the nal product.23 It has
also been shown that, for low-temperature syntheses with
alkanes, various morphologies such as bers and sheets can be
synthesized in 4 h with a retained hexagonal pore structure.22
However, to our knowledge, there are no reports on how to
reduce the reaction time with retained pore structure and particle
size within the rod morphology. The case of dispersed rods is
slightly dierent since the stirring time used is shorter, normally
less than 10 min, to avoid agglomeration of the particles. Instead
the material is kept under static conditions (no stirring), usually
for 20 h,10,14,15 prior to the hydrothermal treatment.
Here we show that it is possible to synthesize monodispersed
SBA-15 rods in the presence of heptane and NH4F in a lowtemperature synthesis. The length of the rods can be varied from
400 to 700 nm by changing the HCl concentration. Furthermore,
the pore size can be tuned between 11 and 17 nm by changing the
hydrothermal treatment time and temperature at a given HCl
concentration. The nal material is formed within 1 h, which is a
Received: December 7, 2010
Revised:
January 31, 2011
Published: March 17, 2011
4994

dx.doi.org/10.1021/la104864d | Langmuir 2011, 27, 49944999

Langmuir

ARTICLE

Table 1. Synthesis Conditions and Physisorption Data for the Synthesized Materials
sample [HCl] (mol/L) aging temp (C) aging time (h) static time (h) specic surface area (m2/g) pore size (nm) unit cell param (nm) total pore vol (cm3/g)

1.37

100

24

560

15.6

15.0

1.34

1.68

100

24

525

15.1

14.4

1.22

1.75

100

24

559

15.4

14.6

1.20

1.83

100

24

542

15.1

14.4

1.20

1.90

100

24

538

13.7

14.2

1.13

1.98

100

24

498

13.7

14.1

1.05

t6

1.75

100a

776

11.2

13.9

1.11

t24
t48

1.75
1.75

100a
100a

24
48

3
3

529
577

14.1
15.4

14.6
14.9

1.06
1.17

t120

1.75

100a

120

438

15.8

14.9

1.12

T80

1.75

80a

24

778

13.1

14.1

1.31

T130

1.75

130a

24

513

16.1

14.9

1.21

T130-t120

1.75

130a

120

410

16.8

15.4

1.12

Stat0

1.75

100

24

0 min

502

16.6

15.1b

1.35

Stat5

1.75

100

24

5 min

562

17.3

15.1

1.43

Stat10
Stat30

1.75
1.75

100
100

24
24

10 min
30 min

560
507

16.7
15.6

14.9
14.6

1.35
1.18

Stat60

1.75

100

24

60 min

545

15.0

14.6

1.17

Stat180

1.75

100

24

180 min

522

15.2

14.6

1.20

Hydrothermal treatment in an autoclave. b Calculated from the (100) diraction peak only.

reduction in reaction time with 19 h compared to a regular


synthesis of dispersed rods.

EXPERIMENTAL SECTION
Synthesis. Hydrochloric acid (purity g37%, puriss. p.a., Fluka, ACS
reagent, fuming), P123 (Aldrich), ammonium fluoride (purity g98.0%,
puriss. p.a., ACS reagent, Fluka), TEOS (reagent grade, 98%, Aldrich),
and heptane (99%, ReagentPlus, Sigma-Aldrich) were used as received.
In a typical synthesis 2.4 g of P123 and 0.028 g of NH4F were
dissolved in 80 mL of HCl solution. The HCl concentration in the
solution was varied between 1.37 and 1.83 M. The mixture was stirred at
20 C until the polymer was dissolved. A 17 mL volume of heptane was
premixed with 5.5 mL of TEOS and then the resulting mixture added to
the micellar solution. The synthesis was kept under vigorous stirring for
4 min and then under static conditions for 04 h. After the reaction the
solution was transferred to an autoclave for hydrothermal treatment at
100 C for 24 h. During the hydrothermal treatment study, the time and
temperature were varied between 6 and 120 h and 80 and 130 C,
respectively. The material was then ltered and washed with distilled
water and dried at 100 C overnight. Finally, the material was calcinated
at 550 C for 5 h. The samples and their synthesis conditions are
described in Table 1.
Characterization. Scanning electron microscopy (SEM) was
performed with a Leo 1550 Gemini scanning electron microscope
operated at 3 kV and a working distance of 35 mm. Nitrogen sorption
isotherms were obtained with a Micromeritics ASAP 2020 at 196 C
with samples outgassed at 300 C for 5 h. The pore size distribution was
calculated from the adsorption isotherm using the KrukJaroniecSayari
method24 and the BrunauerEmmettTeller (BET) surface area from the
relative pressure of 0.060.17. The total pore volume was estimated at
P0/P = 0.975. Transmission electron microscopy (TEM) was performed
with an FEI Tecnai G2 TF 20 UT microscope operated at 200 kV. TEM
samples were prepared by dispersing the product in acetone and
depositing it on a hollow carbon grid. X-ray diffraction (XRD) was
carried out with Cu KR radiation on a Kratky compact small-angle

Figure 1. SEM micrographs of particles synthesized with HCl concentrations of (a) 1.37 M, (b) 1.68 M, (c) 1.75 M, (d) 1.83 M, (e) 1.90 M,
and (f) 1.98 M.
system equipped with a 1024-channel detector. The air scatter was
reduced by using evacuated flight tubes.

RESULTS
Particle Shape and Separation. The particle shape varies
with the HCl concentration as can be seen in the SEM micrographs in Figure 1. In Figure 1a the particles are agglomerated
and 400 nm long and 400 nm wide. Increasing the HCl
concentration to 1.681.75 M (Figure 1b,c) results in separated
4995

dx.doi.org/10.1021/la104864d |Langmuir 2011, 27, 49944999

Langmuir

Figure 2. TEM micrographs of materials synthesized with 1.37 M (a,


b), 1.83 M (c, d), and 1.98 M (e) HCl.

Figure 3. Physisorption isotherms and pore size distributions calculated with the KJS method as a function of the HCl concentration in the
synthesis.

particles. The particles are 400 nm long and have a diameter of


150200 nm. The particles have an elliptic, spindle-like shape,
and the pores in the particle center are longer than the pores
closer to the mantle surface; see the inset in Figure 1b. Increasing
the HCl concentration further elongates the particles to
600 nm, Figure 1d. Higher HCl concentrations (1.901.98
M) yield 500700 nm long and 100 nm wide rodlike particles;
see Figure 1e,f. In these particles, all pores are equally long; see
the inset in Figure 1f.
TEM micrographs, Figure 2, show that the pores are running
along the long axis of the particles and are arranged in a hexagonal
order for both the spindle- and rod-shaped particles. For sample

ARTICLE

Figure 4. SEM and TEM micrographs of materials synthesized for (a,


b) 0 min, (c, d) 5 min, and (e) 60 min static times.

A, synthesized with the lowest HCl concentration, there is a


mixture between small particles with ordered pores and larger
particles with no apparent pore order; see Figure 2a,b. The
physisorption isotherms for all morphologies, Figure 3, are all
type IV isotherms with type 1 hysteresis loops. These isotherms
are typical for SBA-15 with its cylindrical, hexagonally ordered
pores. The pore size distributions for the materials aged at 100 C
for 24 h are narrow and reveal a pore size of 1415 nm; see
Table 1. In Table 1, the unit cell parameters from all samples are
also included. The unit cell reported is the average of the
calculated unit cells from the (100), (110), and (200) diraction peaks.
Formation Time. The static time needed to complete the
formation of particle morphology and pore structure was studied
with SEM and TEM; see Figure 4. It is clear that the static time
prior to hydrothermal treatment is important for the formation of
particles and ordered pores. When the static time is 0 min, the
material consists of small spheres and elongated shapes; see
Figure 4a,b, where the elongated particles have ordered porosity
while the spheres are mostly empty shells. When the static time is
increased to 5 min, the material has the correct rod morphology
and a completely ordered pore structure; see Figure 4c. This
morphology is retained throughout the synthesis; compare with
Figure 4e.
Physisorption isotherms, X-ray diractograms, unit cells, and
pore sizes for materials synthesized with various static times are
presented in Figure 5. It is clear from Figure 5a,b that there is a
mixture of cylindrical and spherical micelles in the material when
no static time is used. The leftmost peak in the X-ray diractogram (for Stat0) is assigned to spherical micelles, while the
second peak, (100), originates from the cylindrical micelles,
starting to form the hexagonal structure. After 5 min of static
time, the material is ordered, which is also seen in the X-ray
diractograms (Figure 5b). As seen in Figure 5c, the pore size
and unit cell parameters decrease with increasing static time until
4996

dx.doi.org/10.1021/la104864d |Langmuir 2011, 27, 49944999

Langmuir

ARTICLE

Figure 6. Pore size distributions for materials synthesized in an autoclave using dierent hydrothermal treatment conditions: (a) temperature and (b) time.

the nal pore size is reached. The pore volume is small for Stat0
but then increases as soon as the ordered pore structure is
present, Stat5 and above. The pore size and unit cell parameter
decrease with increasing static time until the nal value is reached
at Stat60. Samples synthesized with a 60 min static time are
similar to those synthesized with longer times, e.g., 180 min.
Hydrothermal Treatment. The effects of variations in the
hydrothermal treatment time and temperature on the pore size
are seen in Figure 6. All physisorption data are presented in
Table 1. It is clear that increasing the temperature and/or the
time of the hydrothermal treatment yields larger pores and a
decreased specific surface area of the materials.

DISCUSSION

Figure 5. (a) Nitrogen sorption isotherms, (b) X-ray diractograms,


and (c) pore sizes and unit cell parameters for samples synthesized with
various condensation times.

Particle Shape and Separation. The synthesis of SBA-15


rods has been performed by varying the stirring time and HCl
concentration. It has previously been shown that decreased
stirring time is essential for producing monodispersed SBA-15
rods.10 In our work, the separation of particles was only possible
at higher concentrations of HCl, compared to other low-temperature syntheses with heptane and NH4F,21,22 even when the
stirring time was decreased to 5 min. The size and shape of the
particles synthesized with the lowest HCl concentration are
identical to the size and shape of the crystallites building up
the morphologies previously reported for this system.21,22 The
reason for the need of a higher HCl concentration to completely
4997

dx.doi.org/10.1021/la104864d |Langmuir 2011, 27, 49944999

Langmuir
separate the particles originates in the hydrolysis and condensation rates of TEOS and passivation of hydroxyl groups on the
particle surfaces. The hydrolysis and condensation rates of TEOS
are pH dependent, and for pH < 2 both increase with decreasing
pH (increasing HCl concentration).25 Studies of the formation
of SBA-15 have shown that spherical micelles become elongated
to cylinders upon addition of the silica precursor.26,27 These
micelles then attach together and form hexagonally ordered
clusters which then form the particles observed here. The
attachment occurs through linking of micelles by hydroxyl
groups during condensation of the silica species, and with time
these walls become more dense.27 It has previously been shown
that a low pH favors the fiber morphology while a higher pH
favors side by side anchoring of silicate micelles during the
formation. This has been explained by variations of protonation
of the silicates and electrical double-layer repulsion between
colloidal particles when the pH is altered.28,29
The particle length increases with the HCl concentration; cf.
Figure 1. This is in contradiction to what has previously been
reported by Wang et al.,15 who had glycerol present in the
synthesis, aecting the bridging between silica precursors and
thereby their hydrolysis rate. In that study it was stated that
variation in the particle length was due to a combination of high
concentrations of acid, yielding a larger number of seeds from
which particles can be formed, and the increased passivation of
the basal planes, inhibiting further elongation. In our case, the
elongation is rather related to the concentration of HCl in
combination with NH4F and thereby the hydrolysis rate of the
silica precursor. Fluoride ion, F, is a well-known catalyst for
hydrolysis and condensation of TEOS, but the concentration of
F in the solution will be aected by the HCl concentration due
to the fact that some of the H ions will form HF with a fraction
of the F ions. When F ions are in the solution, the silica
oligomers form networks that cannot penetrate deep into the
poly(ethylene oxide) (PEO) shell.30 This yields less exibility
when the silicated micelles agglomerate to form the hexagonal
structure; the walls will be denser earlier and can lock the
formation from evolving further. Increasing the HCl concentration will remove some of the F catalysts, and smaller oligomers
will form the walls, making the silicated micelles more exible.
The micelles can then become elongated to form the hexagonal
structure. At the same time, a decreased pH increases the
hydrolysis and condensation rate of the silica precursor, which
increases the formation rate of the particles. As seen in studies of
the formation of SBA-15, the micelles become more elongated
and narrower with increasing time.26 This is consistent with our
observations; when the HCl concentration is increased, the
particles become more elongated and the pore size is decreased.
Hence, it is the competition between the eect of the decrease in
both the pH and F concentration that yields the variation in
morphology. For the lowest HCl concentration, sample A, the
particles are attached to each other and the pores are not as
ordered. A longer stirring time, e.g., 4 h, is needed to form an
ordered pore structure for this synthesis composition, and in this
case the ber morphology will be formed.22 In the case of
insucient stirring time, the micelles are not elongated and
can therefore not form the ordered structure. Instead the ongoing condensation quenches the material in its disordered state,
and particles are agglomerated and attached to each other with
time. Similar quenching of the structure is gained for higher HCl
concentrations as well. For sample B, there is an ordered pore
structure, but the pore lengths vary within each particle. The HCl

ARTICLE

concentration for this sample is sucient for the formation of


cylindrical micelles that attach in a hexagonal order. However,
some micelles are not as rapidly elongated as others, due to
uctuations in the synthesis, and these are attached on the mantle
surface prior to complete elongation. By increasing the HCl
concentration, the formation rate increases and all micelles reach
similar lengths prior to the end of stirring, samples E and F. The
elongation of all micelles and the eect of the HCl concentration
on the particle morphology are seen in the insets in Figure 1b,d,f.
Formation Time. The static time between stirring and hydrothermal treatment has been studied with the purpose of reducing
the reaction time. The minimum time needed for the synthesis
was defined as the time when the pore size, surface area, and
morphology were similar to those of the material synthesized
with a 3 h static time, sample C. It has previously been shown that
SBA-15 is completely formed after 2 h when normal synthesis
conditions, including continuous stirring, are used.23 Other
morphologies such as sheets with large pores are synthesized
with a 4 h reaction time.22 The data presented here show that the
formation of the rods with large pores is more rapid and the
reaction time can be decreased to 1 h.
If the hydrothermal treatment is started directly after the
solution is stirred for 4 min (Stat0), only a few ordered cylindrical
pores are present in the nal product. When 5 min of static time is
used, the rodlike particles are formed and the hexagonally
ordered cylindrical pores are present. This morphology is then
retained, independent of the length of the static time. The pore
size, on the other hand, decreases with increasing static time, and
a stable value is only present for static times above 60 min. The
decrease is due to condensation of silica in the walls surrounding
the micelles. Studies of the formation of SBA-15 reveal a
decreased diameter of the cylindrical micelles during the formation process.26,27,31,32 In our study, although the hexagonally
ordered structure is formed after 5 min, the silica condensation
and densication of the walls continue for 1 h at 20 C. The
condensation is also present during the hydrothermal treatment,
but at these elevated temperatures the condensation rate is
increased.33 Hence, rapid condensation of silica species during
hydrothermal treatment eectively quenches the morphology
and pore structure and stops further evolution.
Pore Size. Additions of heptane and NH4F in a low-temperature synthesis yield larger pores compared to the original
SBA-15 synthesis. Both NH4 and F are regarded to act as
salting-out ions; i.e., they reduce the solubility of P123 in the
aqueous solution. At this temperature, the P123 concentration
is below the critical micelle concentration (cmc) and micelles
will not form without salt additions. Furthermore, heptane
reduces the hydration of the hydrophobic poly(propylene
oxide) (PPO) chains. Hence, micelles can form at this low
temperature. Heptane is also located at the micelle core acting
as a swelling agent and can be used to control the pore size.22 In
summary, the combination of low temperature, NH4F, and
heptane used in this study yields pores that are significantly
larger than for the original synthesis of SBA-15. Also, F has
been shown to be a nucleophilic catalyst and accelerates the
hydrolysis and condensation rates of TEOS. The increased
hydrolysis and condensation rates due to the catalyst cause the
silica oligomers to form networks that cannot penetrate deep
into the PEO shell since they are more hydrophobic and less
flexible. This increases the nonsilicated region in the PEO shell;
hence, the wall thickness decreases, and the pore size is
increased.30,34
4998

dx.doi.org/10.1021/la104864d |Langmuir 2011, 27, 49944999

Langmuir
The decrease in pore size with increasing HCl concentration is
probably due to two eects: increased formation rate with
increasing HCl concentration and a competition between H
and F with increasing HCl concentration. As discussed previously, the increased hydrolysis rate of the silica precursor due to
increased HCl concentrations in combination with the short
stirring time yields micelles that are more elongated and narrower. Furthermore, the balance among H, F, and HF in the
solution is shifted toward HF when the H concentration is
increased. Hence, the amount of F in the solution is reduced,
the condensation promoter eect of these ions is decreased, the
silica oligomers can penetrate deeper into the PEO shell, and the
pore size is decreased.
Hydrothermal treatment is a known method for controlling the
pore size at the angstrom level. When hexane has been used as a
swelling agent, an increase of the hydrothermal treatment time and
temperature to 130 C for 5 days increases the pore size to 18.2 nm
compared with 13.9 nm with 100 C for 1 day.35 When increasing
temperature, the hydrophilicity of the PEO chains decreases. The
chains then redraw from the silica network into the hydrophobic core
of the micelles.36 This leads to an increased pore size, decreased wall
thickness, and reduced specic surface area of the material. Hence, the
pore size can be precisely tuned between 11.2 and 16.8 nm, according
to the KJS method, by alterations in the hydrothermal treatment time
and temperature. It should be noted though that the KJS method
overestimates the pore size for pores larger than 12 nm;37 e.g., it has
previously been seen that a pore size of 19.9 nm calculated with the
KJS method is from TEM micrographs estimated to be 18 nm.21 This
is clearly also the case here, where the unit cell parameter is smaller
than the pore size for all samples; see Table 1.

CONCLUSION
We show that it is possible to synthesize separate SBA-15
particles in a low-temperature synthesis with additions of heptane and NH4F by decreasing the stirring time during the
reaction. Also, the reaction time can be decreased from 20 to 1
h without losing the quality of the material. The length and shape
of the particles can be controlled by varying the HCl concentration, and a mechanism for these variations is suggested. The pore
size can be tuned between 11.3 and 16.8 nm by changing the
hydrothermal treatment time and temperature. This work has
resulted in SBA-15 rods with the largest pore size reported for
this morphology.
AUTHOR INFORMATION
Corresponding Author

*E-mail: emmjo@ifm.liu.se. Phone: 46 (0)13 282543. Fax:


46 (0)13 288918.

ACKNOWLEDGMENT
Dr. Jessica Rosenholm, bo Akademi, is greatly acknowledged
for help with the XRD measurements. The Swedish Research
Council (VR) is acknowledged for nancial support.
REFERENCES
(1) Lee, J.; Cho, D.; Shim, W.; Moon, H. Korean J. Chem. Eng. 2004,
21, 246251.
.; Boissiere, C. C. R. Chim. 2005, 8, 579596.
(2) Prouzet, E
(3) Chao, K.; Liu, P.; Huang, K. C. R. Chim. 2005, 8, 727739.

ARTICLE

(4) Giraldo, L. F.; Lopez, B. L.; Perez, L.; Urrego, S.; Sierra, L.; Mesa,
M. Macromol. Symp. 2007, 258, 129141.
(5) Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B. F.; Stucky, G. D. J. Am.
Chem. Soc. 1998, 120, 60246036.
(6) Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G. H.;
Chmelka, B. F.; Stucky, G. D. Science 1998, 279, 548552.
(7) Schmidt-Winkel, P.; Yang, P.; Margolese, D. I.; Chmelka, B. F.;
Stucky, G. D. Adv. Mater. 1999, 11, 303307.
(8) Ma, Y.; Qi, L.; Ma, J.; Wu, Y.; Liu, O.; Cheng, H. Colloids Surf., A
2003, 229, 18.
(9) Chen, S.; Tang, C.; Chuang, W.; Lee, J.; Tsai, Y.; Chan, J. C. C.;
Lin, C.; Liu, Y.; Cheng, S. Chem. Mater. 2008, 20, 39063916.
(10) Sayari, A.; Han, B.-H.; Yang, Y. J. Am. Chem. Soc. 2004,
126, 1434814349.
(11) Fan, J.; Lei, J.; Wang, L.; Yu, C.; Tu, B.; Zhao, D. Chem.
Commun. 2003, 21402141.
(12) Lei, J.; Fan, J.; Yu, C.; Zhang, L.; Jiang, S.; Tu, B.; Zhao, D.
Microporous Mesoporous Mater. 2004, 73, 121128.
(13) Sun, J.; Zhang, H.; Tian, R.; Ding, M.; Bao, X.; Su, D. S.; Zou, H.
Chem. Commun. 2006, 13221324.
(14) Yu, C.; Fan, J.; Tian, B.; Zhao, D.; Stucky, G. D. Adv. Mater.
2002, 14, 1742.
(15) Wang, Y.; Zhang, F.; Wang, Y.; Ren, J.; Li, C.; Liu, X.; Guo, Y.;
Guo, Y.; Lu, G. Mater. Chem. Phys. 2009, 115, 649655.
(16) Kosuge, K.; Sato, T.; Kikukawa, N.; Takemori, M. Chem. Mater.
2004, 16, 899905.
(17) Fulvio, P. F.; Pikus, S.; Jaroniec, M. J. Mater. Chem. 2005,
15, 50495053.
(18) Lettow, J. S.; Han, Y. J.; Schmidt-Winkel, P.; Yang, P.; Zhao, D.;
Stucky, G. D.; Ying, J. Y. Langmuir 2000, 16, 82918295.
(19) Sun, J.; Zhang, H.; Ding, M.; Chen, Y.; Bao, X.; Klein-Homan,
A.; Pfander, N.; Su, D. S. Chem. Commun. 2005, 53435345.
(20) Zhang, H.; Sun, J.; Ma, D.; Weinberg, G.; Su, D. S.; Bao, X.
J. Phys. Chem. B 2006, 110, 2590825915.
(21) Johansson, E. M.; C
ordoba, J. M.; Oden, M. Mater. Lett. 2009,
63, 21292131.
(22) Johansson, E. M.; Cordoba, J. M.; Oden, M. Microporous
Mesoporous Mater. 2010, 133, 6674.
(23) Fulvio, P. F.; Pikus, S.; Jaroniec, M. J. Colloid Interface Sci. 2005,
287, 717720.
(24) Kruk, M.; Jaroniec, M.; Sayari, A. Langmuir 1997, 13, 62676273.
(25) Cihlar, J. Colloids Surf., A 1993, 70, 239251.
(26) Ruthstein, S.; Schmidt, J.; Kesselman, E.; Telmon, Y.; Goldfarb,
D. J. Am. Chem. Soc. 2006, 128, 33663374.
(27) Khodakov, A. Y.; Zholobenko, V. L.; Imperor-Clerc, M.;
Durand, D. J. Phys. Chem. B 2005, 109, 2278022790.
(28) Yang, S. M.; Yang, H.; Coombs, N.; Sokolov, I.; Kresge, C. T.;
Ozin, G. A. Adv. Mater. 1999, 11, 5255.
(29) Ozin, G. A.; Kresge, C. T.; Yang, H. In Nucleation, Growth and
Form of Mesoporous Silica: Role of Defects and a Language of Shape;
Bonneviot, L., Beland, F., Danumah, C., Giasson, S., Kaliaguine, S., Eds.;
Studies in Surface Science and Catalysis, Vol. 117; Elsevier: New York,
1998; pp 119127.
(30) Boissiere, C.; Martines, M. A. U.; Tokumoto, M.; Larbot, A.;
Prouzet, E. Chem. Mater. 2003, 15, 509515.
(31) Flodstr
om, K.; Teixeira, C. V.; Amenitsch, H.; Alfredsson, V.;
Linden, M. Langmuir 2004, 20, 48854891.
(32) Flodstr
om, K.; Wennerstrom, H.; Alfredsson, V. Langmuir
2004, 20, 680688.
(33) Colby, M. W.; Osaka, A.; Mackenzie, J. D. J. Non-Cryst. Solids
1986, 82, 3741.
(34) Boissiere, C.; van, d. L.; El Mansouri, A.; Larbot, A.; Prouzet, E.
Chem. Commun. 1999, 20472048.
(35) Kruk, M.; Cao, L. Langmuir 2007, 23, 72477254.
(36) Galarneau, A.; Cambon, H.; Di Renzo, F.; Fajula, F. Langmuir
2001, 17, 83288335.
(37) Jaroniec, M.; Solocyoc, L. A. Langmuir 2006, 22, 67576760.
4999

dx.doi.org/10.1021/la104864d |Langmuir 2011, 27, 49944999

Das könnte Ihnen auch gefallen