Sie sind auf Seite 1von 5

By Xuchuan Jiang, Thurston Herricks, and Younan Xia*

Spherical colloids have recently been exploited by many research groups as building blocks to fabricate photonic crystals
via self-assembly.[1] The majority of these studies, however,
have been limited to polymer latexes and silica spheres because of the easiness in processing these materials as spherical
colloids with truly monodisperse sizes and in copious quantities. Only a few demonstrations involved the synthesis and
crystallization of spherical colloids made of compound semiconductors such as CdS, ZnS, and ZnS (deposited on polystyrene spheres).[2] Photonic band structures of these new systems have been shown to exhibit features different from those
of conventional opals due to their higher refractive indices
relative to polystyrene (~ 1.6) or silica (~ 1.5).[3] Among various inorganic semiconductors, titania has long been considered as an ideal candidate for generating photonic crystals
due to its low absorption in the visible and near-infrared regions and its relatively high refractive indices (2.4 for anatase
and 2.9 for rutile).[3] Unfortunately, no one has been able to
prepare titania as monodispersed spherical colloids with size
variations within 5 %.
Because of its technological importance in various industrial applications (e.g., as pigments or paper whiteners,[4]
photocatalysts,[5] and optical coatings[6]), a large number of
chemical methods have been developed for generating colloidal particles of titania. In industry, titania particles are often
synthesized by digesting the ore ilmenite with sulfuric acid,
followed by thermal hydrolysis of the titanium(IV) ions in a
highly acidic solution and eventually dehydration of the titanium(IV) hydrous oxides. The resulting particles were often irregular in shape and exhibited broad distributions in size.
Like many other ceramic oxides, titania has also been synthesized as spherical colloids by controlling the hydrolysis and
condensation of an appropriate precursor.[7] For example,
Matijevic and co-workers have developed a procedure to generate titania spherical colloids with a relatively narrow distribution in size (r = 1525 %) by reacting water vapor with a
titania precursor contained in liquid aerosols.[8] They have
also demonstrated that hydrolysis of TiCl4 in the presence of
sulfate ions at elevated temperatures could lead to the formation of stable aqueous dispersions of titania spheres with di-

[*] Prof. Y. Xia, Dr. X. Jiang


Department of Chemistry, University of Washington
Seattle, WA 98195 (USA)
E-mail: xia@chem.washington.edu
T. Herricks
Department of Materials Science and Engineering
University of Washington
Seattle, WA 98195 (USA)

[**] This work has been supported in part by the STC Program funded by the
National Science Foundation (NSF) under Agreement Number DMR0120967, a Career Award from the NSF (DMR-9983893), and a Fellowship from the David and Lucile Packard Foundation. Y. X. is a Camille
Dreyfus Teacher Scholar and an Alfred P. Sloan Research Fellow.

Adv. Mater. 2003, 15, No. 14, July 17

DOI: 10.1002/adma.200305105

ameters ranging from 1 to 4 lm.[9] Barringer, Bowen, Ring,


and their co-workers have prepared titania spheres 300
700 nm in diameter by controlling the hydrolysis of titanium
tetraethoxide in dilute alcoholic solutions.[10] They have also
prepared fine particles of titania with relatively uniform sizes
by adding hydroxypropyl cellulose (a polymeric surface stabilizer) to the solgel solution.[11] Because the titania spheres
synthesized using these methods were not truly monodispersed (often with size variations > 10 %), it has been difficult
to crystallize them into long-range ordered lattices.
It has been suspected that the conventional solgel methods
could not be used to generate monodispersed spherical colloids of titania because the hydrolysis rates of these precursors
were too fast, and thus the nucleation and growth were never
separated into two steps. Because the concentration of titania
monomer was always higher than the critical value, nucleation
events could continuously occur in the entire solgel process,
leading to the formation of polydispersed samples. By switching to solgel precursors with significantly lower hydrolysis
rates, one should be able to obtain titania spherical colloids
with narrower distributions in size. For example, spherical colloids with improved monodispersity in size have been synthesized by using a precursor, Ti(OPr)3(acac), derived from the
modification of Ti(OPr)4 with acetylacetone (acac).[12] In the
present work, we found that glycols could serve a class of
reagents to greatly reduce the hydrolysis rates of titanium
alkoxides. When mixed with titanium alkoxide, glycols were
sufficiently reactive to form glycolates or mixed alkoxide/glycolate derivatives (see the reactions below with ethylene glycol as an example):[13]
Ti(OBu)4 + HOCH2CH2OH ?
Ti(OCH2CH2O)(OBu)2 + 2 HOBu

(1)

Ti(OBu)4 + 2 HOCH2CH2OH ?
Ti(OCH2CH2O)2 + 4 HOBu

(2)

Different from titanium alkoxides that are highly susceptible to moisture (white precipitation was immediately formed
when it was exposed to air), the glycolated precursors are
more resistant to hydrolysis, and could be kept in air for several months without observing any precipitation from solution.
When such an ethylene glycolated precursor was poured into
an acetone bath containing a small amount of water (~ 0.3 %),
spherical colloids of the titanium glycolate were obtained in
copious quantities through a homogeneous nucleation and
growth process. The exact function of acetone is yet to be understood, and we suspected that it might act as a catalyst to
speed up the hydrolysis rate of the glycolated precursor. Because the nucleation events could only occur at the beginning
of this process (as a result of fast growth rates for the colloids), the formation of uniform spherical colloids was ensured.
Figure 1 shows the scanning electron microscopy (SEM)
images of colloidal particles that were synthesized with different molar concentrations of Ti(OBu)4-derived glycolate in ac 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1205

COMMUNICATIONS

Monodispersed Spherical Colloids of Titania:


Synthesis, Characterization, and Crystallization**

COMMUNICATIONS

(110)

(211)
(101) (111)

Intensity (a.u.)

(002)
(310) (301)
(220)

(200) (210)

950 C

650 C
(101)

(103) (004)

500 C

rutile

* anatase
*
*

*
o

(200)

*
*

(105)(211)
(204) (116)

350 C
R.T.

10

20

30

40

50

60

70

2 (degree)
Fig. 2. XRD patterns of a colloidal sample showing the formation of various
phases at different temperatures: titanium glycolate (RT); amorphous (350 C);
anatase (500 C); a mixture of anatase and rutile (650 C); and rutile (950 C).

1206

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

110

A
100

Weight Loss (%)

etone. These particles were spherical in shape, and characterized by smooth surfaces. As the precursor concentration was
increased, there was a monatomic increase in the diameter. It
was difficult to obtain a straightforward correlation between
the diameter of spherical colloids and the precursor concentration because the number of nuclei seemed to have a complex, nonlinear dependence on the precursor concentration.
We could easily tune the diameter of these colloids from 200
to 500 nm by varying the precursor concentration in the range
from 0.68 to 1.2 mM (calculated as the final concentration of
the precursor in acetone). Without changing the current procedure, use of precursor solutions with concentrations outside
this region often led to the formation of polydispersed samples. It's worth noting that even with the spherical colloids
(200500 in diameter) that we have synthesized, it should be
possible to fabricate colloidal crystals with stop bands spanning across the spectral region from visible to near infrared
(NIR).[14]
The spherical colloids could be converted from the titanium
glycolate to pure anatase and then to rutile by annealing them
in air at elevated temperatures. Figure 2 shows X-ray diffraction (XRD) patterns taken from a sample of 500 nm spherical
colloids, and after this sample had been annealed for 2 h at
various temperatures. It is clear that the as-synthesized particles were amorphous in structure with essentially no diffraction peaks. The samples obtained by annealing at 350, 500, and
950 C were assigned to the amorphous, anatase (JCPDS File
No. 21-1272) and rutile phase (JCPDS File No. 21-1276) of
titania. From the XRD pattern, the phase transition from anatase to rutile seemed to occur at temperatures around 650 C,
where diffraction peaks of both phases were detected in the
sample. These observations were consistent with previously reported results based on conventional solgel precursors.[15]

The compositional and structural changes associated with


such a thermal annealing process were also characterized
using thermogravimetric analysis (TGA) and Fourier transform infrared spectroscopy (FTIR). The samples were prepared by drying in a vacuum oven at 40 C for 6 h. Figure 3A
shows the TGA curve recorded under a flow of nitrogen gas,
indicating a two-step pattern for weight loss in the tempera-

90
80
70
60
100

200

300

400

500

Temperature ( C)

500 C

Transmittance (a. u.)

Fig. 1. SEM images of spherical colloids of titania glycolates that were prepared
with different molar concentrations of Ti(OBu)4 in acetone: A) 1.21, B) 1.03,
C) 0.86, and D) 0.68 mM. The mean diameters (D) of these colloids were
500 13.8 nm (evaluation of 39 particles); 400 5.2 nm (83 particles);
320 9.7 nm (125 particles); and 200 11.6 nm (394 particles), respectively.
The uncertainty was calculated as variance using a computer program (Origin).

350 C

*
o

200 C

*
o

100 C

*
*
4000

3000

2000

1000

R.T.

-1

Wavenumber (cm )
Fig. 3. A) TGA curve recorded from colloidal particles of titanium glycolate.
The changes in weight correspond to desorption of physically adsorbed water
(25100 C), and desorption of chemically adsorbed water and ethylene glycol
units (150350 C). B) FTIR spectra taken from the spherical colloids and after
it had been annealed at various temperatures from RT to 500 C. Asterisk and
solid circle denote the TiO stretching mode in the glycolate precursor and the
amorphous phase of titania (at ~ 640 cm1), and in the anatase phase (at
~ 465 cm1), respectively.

http://www.advmat.de

Adv. Mater. 2003, 15, No. 14, July 17

Fig. 4. A,B) SEM and C,D) TEM images of spherical colloids of titania glycolate after they had been heated at A,C) 500 and B,D) 950 C for 2 h.
E,F) Electron diffraction patterns recorded from these two samples. Each pattern could be indexed to the pure anatase and rutile phase, respectively.

Adv. Mater. 2003, 15, No. 14, July 17

http://www.advmat.de

(SEM) (A,B) and transmission electron microscopy (TEM)


(C,D) images of a sample (~ 500 nm in diameter) after it had
been annealed at 500 and 950 C for 2 h. The corresponding
diameters of colloids were reduced by ~ 14 % and ~ 17 %, respectively. These images suggest that the spherical morphology of these colloidal particles were essentially preserved in
the annealing process. Some of the particles might be fused
together (as indicated by arrows) to generate larger aggregates. Another obvious change is that the surfaces of these
colloidal particles were significantly roughened as a result of
crystallization. A closer inspection of the SEM and TEM images indicates that each colloidal particle is composed of crystallites 2030 nm in dimension after the sample had been annealed at 950 C for 2 h. Figures 4E,F show the electron
diffraction patterns recorded from these two samples. The
rings shown in Figure 4E were indexed to diffraction from the
(101), (004), (200), (105), (204), (116), and (215) planes of
anatase, while the rings in Figure 4F could be ascribed to diffraction from the (110), (101), (111), (211), (310), and (301)
planes of rutile. These structural and phase assignments were
consistent with the XRD patterns depicted in Figure 2.
Because of their spherical shape and uniform size, the colloids (made of titanium glycolate) synthesized using this
method could directly serve as building blocks to form 3D
opaline lattices through self-assembly. Figures 5A,B show the
SEM images of an opaline lattice that was crystallized from
spherical colloids of glycolated precursor that were ~ 280 nm
in diameter. This crystalline lattice had a face-centered cubic
(fcc) structure, with its (111) plane oriented parallel to the surface of the supporting substrate. Figure 5C shows an oblique
SEM image of this crystalline lattice after it had been annealed at 500 C for 2 h. Although the shape of these spherical colloids was not significantly deformed in the annealing

Fig. 5. SEM images of a three-dimensional crystal of spherical colloids of titania


glycolate before (A,B) and after (C) it had been heated at 500 C for 2 h.
D) UV-vis reflectance spectra taken from this crystal before (solid line, TiO2
glycolate precursor) and after (dash line, anatase) thermal annealing. The light
was at normal incidence for both measurements.

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1207

COMMUNICATIONS

ture ranges of 25100 C and 150350 C. The first weight loss
corresponds to desorption of physically adsorbed water molecules, and the second one is associated with the removal of
ethylene glycol units and the degradation of organic groups
contained in the precursor particles.[16] Typically, we observed
a weight loss of ~ 3.1 % for physically adsorbed water and
~ 34.9 % for the ethylene glycol units chemically bonded to titanium. Figure 3B shows FTIR spectra taken from samples
that had been annealed at various temperatures from room
temperature (RT) to 500 C. As the temperature was increased, the peaks corresponding to physically absorbed water
(the OH stretching mode at ~ 3400 cm1 and the OH bending mode at ~ 1640 cm1) and the peaks corresponding to
the ethylene glycol units (the OH stretching mode at
~ 3350 cm1) were gradually reduced in intensity up to 350 C,
and then they all disappeared except the TiO stretching
mode at ~ 640 cm1. This TiO stretching band was shifted to
~ 465 cm1 once the amorphous structure had been rearranged
into the anatase phase of titania at 500 C. These observations
were consistent with the literature.[17]
Electron microscopy and diffraction studies were performed to characterize the morphology, microstructure, and
crystallinity of the spherical colloids obtained through thermal
annealing. Figure 4 shows scanning electron microscopy

COMMUNICATIONS

process, their average size was reduced by ~ 16 %. This large


volume shrinkage resulted in the formation of some additional cracks in the annealed sample. Figure 5D shows the reflection spectra (at normal incidence and detection) taken
from this crystalline sample before and after thermal annealing. The blue shift (from ~ 756 to ~ 687 nm) observed for the
stop band was mainly caused by the reduction in lattice constant (due to volume shrinkage). By fitting peak positions
with the Bragg diffraction equation (the filling fraction was
assumed to be 74 % for both samples), a refractive index of
1.88 and 2.06 was calculated for the precursor and anatase
particles, respectively. These two numbers are consistent with
the values documented in the literature. For example, Khoo
and co-workers reported a refractive index of ~ 2 for an array
of titania microstructures fabricated by patterning a solgel
film with interference photolithography.[18] Wijnhoven and
Vos have studied photonic bandgap properties of inverse
opals made of titania (derived from a solgel precursor), and
a refractive index of ~ 1.89 was measured for their samples.[19]
Comparing the refractive index of this sample with that of
anatase single crystal (2.4),[3] the relatively lower values associated with the solgel-derived samples could be ascribed to
the air pores trapped in these polycrystalline materials.
In summary, monodispersed spherical colloids of titania
have been synthesized in large quantities using glycolated precursors. The hydrolysis rate of a conventional solgel precursor could be significantly slowed and thus better controlled by
forming a coordination complex with ethylene glycol. Spherical colloids of titanium glycolates with uniform diameters in
the 200500 nm range could be conveniently synthesized
using this method. These spherical colloids could be readily
assembled into crystalline lattices with qualities sufficiently
good for optical characterization. We have also demonstrated
that these spherical colloids could be converted from titanium
glycolate to anatase and then to rutile by annealing the samples at elevated temperatures in air. We believe the availability of these monodispersed spherical colloids with high refractive indices in copious quantities will enrich our studies of
photonic bandgap crystals associated with colloidal crystals.

Experimental
In a typical synthesis, 0.050 mL tetrabutoxytitanium (TBT, Aldrich) was
added to 10 mL ethylene glycol (EG, Fisher) in a glove box purged with nitrogen gas. The solution was magnetically stirred for 8 h at room temperature, then
taken out from the glove box and immediately poured into an acetone (Fisher,
A949-4, HPLC grade) bath containing a small amount of water (~ 0.3 %) under
vigorous stirring for 10 min. After aging for ~ 30 min, the white precipitate was
harvested by centrifugation, followed by washing with distilled water and ethanol several times to remove EG from the surfaces of the titania glycolate particles. Monodispersed particles were only obtained when the concentration of
water (in acetone) was kept between 0.3 % and 0.5 %. The diameter of these
particles monotonically increased as the concentration of TBT in the acetone
bath was raised. The as-synthesized particles could be redispersed to form
stable suspensions in ethanol and water without introducing any surfactants.
Other tetraalkoxytitaniums, Ti(OR)4 (R = C2H5 and iso-C3H7, both from Aldrich), were also successfully used as precursors in preparing uniform spherical
colloids. The spherical colloids made of titanium glycolates could be dispersed
in water and then crystallized into an opaline lattice in a microfluidic cell constructed by sandwiching a rectangular gasket (~ 20 lm thick, cut out of a Mylar

1208

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

film) between two glass or one glass and one silicon substrates. The detailed
procedure has already been described in a previous publication [20]. After
water had completely evaporated, the top substrate was carefully separated
from the fluidic cell, leaving behind an opaline lattice on the bottom substrate.
This sample could be directly placed in a box oven to anneal the titanium glycolates into anatase at elevated temperatures in air.
We have also tried many other organic solvents (all from Fisher) such as
methanol (A452-1), 1-propanol (A414-500), 1-butanol (A399-500), acetonitrile
(A998-1), ethyl acetate (E195-4), dimethyl sulfoxide (D128-500), and N,N-dimethyl formamide (D119-500). Spherical colloids of titanium glycolate were
formed in all these solvents, but they were not monodispersed in size.
SEM images were taken using a field-emission microscope (Sirion, FEI,
Portland, OR) operated at an acceleration voltage of 5 kV. TEM images and
electron diffraction patterns were taken using a Phillips EM-430 machine operated at 200 kV. The samples were prepared by placing one drop of the alcohol
suspension on a silicon substrate or TEM grid, and letting the solvent evaporate
slowly in a fume hood. XRD patterns were recorded from the particles using a
Philips PW1710 diffractometer (Cu Ka radiation, k = 1.54056 ) at a scanning
rate of 0.02 per second for 2h in the range from 10 to 70. FTIR spectra of the
samples were acquired in air using a 1720 FTIR spectrophotometer with a
0.5 cm1 resolution (Perkin-Elmer, Norwalk, CT). TGA measurements were
performed on TGA-50/50H analyzer (Shimadzu, Kyoto, Japan) with a heating
rate of 20 C min1 under a flow of nitrogen gas. The reflectance spectra were
measured using an optical fiber spectrometer (S2000, Ocean Optics, FL), with a
beam spot of ~ 0.04 mm2 in area.

Received: March 14, 2003


Final version: April 3, 2003

[1] a) A special issue on photonic crystals: Adv. Mater. 2001, 13(6). b) A. Polman, P. Wiltzius, Materials Science Aspects of Photonic Crystals, a special issue in MRS Bull. 2001, 26, 608. c) O. D. Velev, A. M. Lenhoff, Curr.
Opin. Colloid Interface Sci. 2000, 5, 56. d) A. Stein, R. C. Schroden, Curr.
Opin. Solid State Mater. Sci. 2001, 5, 553. e) A. Blanco, E. Chomski,
S. Grabtchak, M. Ibisate, S. John, S. W. Leonard, C. Lopez, F. Meseguer,
H. Miguez, J. P. Mondia, G. A. Ozin, O. Toader, H. M. van Driel, Nature
2000, 405, 437. f) Y. A. Vlasov, X. Z. Bo, J. C. Sturm, D. J. Norris, Nature
2001, 414, 289. g) J. H. Holtz, J. S. W. Holtz, C. H. Munro, S. Asher, Anal.
Chem. 1998, 70, 780. h) J. F. Bertone, P. Jiang, K. S. Hwang, D. M. Mittleman, V. L. Colvin, Phys. Rev. Lett. 1999, 83, 300. i) D. Wang, R. A. Caruso, F. Caruso, Chem. Mater. 2001, 13, 364. j) W. M. Lee, S. A. Prunziski,
P. V. Braun, Adv. Mater. 2002, 14, 271. k) G. A. Ozin, S. M. Yang, Adv.
Funct. Mater. 2001, 11, 95. l) T. Cassagneau, F. Caruso, Adv. Mater. 2002,
14, 34.
[2] a) E. Matijevic, D. Murphy-Wilhelmy, J. Colloid Interface Sci. 1982, 86,
476. b) D. Murphy-Wilhelmy, E. Matijevic, J. Chem. Soc. Faraday Trans.
1984, 80, 563. c) K. P. Velikov, A. van Blaaderen, Langmuir 2001, 17,
4779. d) M. L. Breen, A. D. Dinsmore, R. H. Pink, S. B. Qadri, B. R. Ratna, Langmuir 2001, 17, 903.
[3] a) Polymer Handbook , 4th ed. (Eds: J. Brandrup, E. H. Immergut, E. A.
Grulke, A. Abe, D. R. Bloch), John Wiley and Sons, New York 1999,
Vol. 93. b) CRC Handbook of Chemistry and Physics, 60th ed., Chemical
Rubber Corp., Boca Raton, FL 19791980, B121, B137.
[4] W. P. Hsu, R. C. Yu, E. Matijevic, J. Colloid Interface Sci. 1993, 156, 56.
[5] M. Anpo, T. Shima, S. Kodama, Y. Kubokawa, J. Phys. Chem. 1987, 91,
4305.
[6] G. A. Battiston, R. Gerbasi, M. Porchia, L. Rizzo, Chem. Vap. Deposition
1999, 5, 73.
[7] J. Livage, M. Henry, C. Sanchez, Prog. Solid State Chem. 1988, 18, 259.
[8] E. Matijevic, M. Budnik, L. Meites, J. Colloid Interface Sci. 1977, 61, 302.
[9] M. Visca, E. Matijevic, J. Colloid Interface Sci. 1979, 68, 308.
[10] a) E. A. Barringer, H. K. Bowen, J. Am. Ceram. Soc. 1982, 65, C199.
b) J. H. Jean, T. A. Ring, Langmuir 1986, 2, 251.
[11] J. H. Jean, T. A. Ring, Colloids Surf. 1988, 29, 273.
[12] a) A. Laustic, F. Bahonneau, J. Livage, Chem. Mater. 1989, 1, 248. b) A.
Yamamoto, S. Kambara, J. Am. Chem. Soc. 1957, 79, 4344.
[13] a) R. C. Mehrotra, PhD Thesis, London University 1952. b) D. M. Puri,
PhD Thesis, University of Gorakhpur 1962. c) D. M. Puri, R. C. Mehrotra, Indian J. Chem. 1967, 51, 448.
[14] S. H. Park, Y. Xia, Langmuir 1999, 15, 266.
[15] a) D. Seo, J. Lee, E. Lee, H. Kim, Mater. Lett. 2001, 51, 115. b) W. Ma,
Z. Lu, M. Zhang, Appl. Phys. A 1998, 66, 621. c) Y. Zhang, A. Weidenkaff,
A. Reller, Mater. Lett. 2002, 54, 375.
[16] M. Popa, M. Kakihana, Solid State Ionics 2002, 151, 251.
[17] a) K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds, Wiley, New York 1978, p. 220. b) D. C. Bradley, R. C.
Mehrotra, D. P. Gaur, Metal Alkoxides, Academic Press, New York 1978,

http://www.advmat.de

Adv. Mater. 2003, 15, No. 14, July 17

Carbon Monoliths Possessing a Hierarchical, Fully


Interconnected Porosity**
By Akira Taguchi, Jan-Henrik Smtt, and Mika Lindn*
Nanocasting of carbon replicas of siliceous micro- or mesoporous materials has gained a lot of interest during the last
years. Micro- and mesoporous carbons, respectively, are, depending on their pore size, interesting materials for a wide
range of applications, including hydrogen storage, doublelayer capacitors, molecular separation, and catalysis. Recently,
the synthesis of highly ordered micro- and mesoporous carbons possessing a narrow pore size distribution has been described.[1,2] Here, mesoscopically ordered silica is impregnated
with a carbon precursor, which is subsequently carbonized
under non-oxidizing conditions. Porous carbons are finally obtained through dissolution of the silica framework. In order to
maintain the structural integrity of the thus prepared carbon
matrix, the host matrix should have an interconnected porosity. Thus, suitable zeolites,[3] MCM-48,[2,4] SBA-1,[5] mesocellular foams,[6] SBA-15,[7] and HMS[8] (hexagonal mesoporous
silica) materials, which all possess a three-dimensional (3D)
interconnected porosity, have been found to be suitable template structures. To date, however, most of the carbon materials reported have been obtained as powders, with only a few
exceptions.[9] This fact may limit the applicability of these
materials when macroscopic morphologies, such as chromatographic columns or membrane reactors, are required. The
present communication is, to the best of our knowledge, the
first to describe the preparation of monolithic carbon possessing a hierarchical bimodal meso- and macroporosity. In a
series of papers, Nakanishi et al.[10] have described a solgel
synthesis route to monolithic silica possessing a bimodal, hierarchical meso- and macroporous structure. This type of monolith is now commercially available as a chromatographic column under the brand name Chromolith. The key to the
synthesis is to balance the kinetics of phase separation versus
gelation of the silica under acidic conditions; this can be
achieved by the use of either homo-polymers or block co-

[*] Dr. M. Lindn, J.-H. Smtt


Dept. of Physical Chemistry, bo Akademi University
Porthaninkatu 35, FIN-20500 Turku (Finland)
E-mail: mlinden@abo.fi
Dr. A. Taguchi
Max-Planck Institut fr Kohlenforschung
Kaiser-Wilhelm Platz 1, D-45470 Mlheim an der Ruhr (Germany)

[**] The financial support from the Finnish Academy of Sciences is gratefully
acknowledged (JHS). The authors thank Mr. Axel Dreier, MPI-Kohlenforschung, Mlheim, Germany, for carrying out the TEM measurements.
We also thank Prof. Ferdi Schth for support and valuable discussions.

Adv. Mater. 2003, 15, No. 14, July 17

DOI: 10.1002/adma.200304848

polymers possessing hydrogen-bonding capability. In a recent


communication,[11] we used a similar approach to obtain hierarchical silica possessing fully interconnected meso- and
macropores with a narrow pore size distribution, by using a
mixture of poly(ethylene glycol) (PEG) and alkyltrimethylammonium bromide (CnTAB) as structure-directing agents.
The pore diameter of both the surfactant-templated mesopores and the macropores can be independently adjusted
between 24 nm and 0.530 lm, respectively, by varying the
hydrocarbon chain length of the surfactant, the PEG molecular weight, the CnTAB/PEG/TEOS ratio, the silica/water ratio
and the post-synthesis treatment conditions. The shape of the
monolithic body can be chosen more or less at will, since it is
determined by the shape of the vessel in which the synthesis is
carried out. The 3D interconnected mesoporosity of this
monolithic silica makes it suitable for the preparation of carbon monoliths possessing a hierarchical meso- and macroporous structure, where the macroscopic morphology is the same
as for the starting silica monolith, as shown in Figure 1. The
enhanced accessibility of the mesopores due to the hierarchical and interconnected macro-mesoporous network is thought

Fig. 1. Photograph showing a meso-macroporous silica monolith used as the


mold for carbon (left) and a nano-cast carbon replica (right). The diameter of
the 1 Euro coin is 23 mm.

to greatly ease the adsorption of larger molecules, which has


been shown to be a kinetically unfavorable process.[12] The
prepared porous silica monoliths, typically rods with a diameter of 5 mm and a length of 1 cm, were evacuated for 2 h at
room temperature (RT) and then impregnated by furfuryl alcohol (FA) by incipient wetness (the material is impregnated
with the solution such that all the pores are filled). FA polymerization and carbonization and subsequent removal of the
silica portion were carried out as described in the Experimental section. As an example, a scanning electron microscopy
(SEM) image of one thus-prepared carbon monolith is shown
in Figure 2, together with a corresponding image of the starting silica monolith.[13] The pore sizes are about 3 lm in both
cases with a wall thickness of about 2 lm. It is evident that a
positive carbon replica of the silica monolith is obtained on
the micrometer scale possessing a fully interconnected macroporosity. Thus, carbon monoliths with macropore diameters
ranging from 0.5 to 30 lm with a narrow pore size distribution
can easily be prepared according to the same procedure. No
change in the macroscopic appearance of the carbon was observed upon removal of the silica matrix. The residual silica

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1209

COMMUNICATIONS

p. 119. c) A. J. Maira, J. M. Coronado, V. Augugliaro, K. L. Yeung, J. C.


Conesa, J. Soria, J. Catal. 2001, 202, 413.
[18] A. Shishido, I. B. Diviliansky, I. C. Khoo, T. S. Mayer, S. Nishimura, G. L.
Egan, T. E. Mallouk, Appl. Phys. Lett. 2001, 79, 3332.
[19] J. E. G. J. Wijnhoven, W. L. Vos, Science 1998, 281, 802.
[20] Y. Lu, Y. D. Yin, B. Gates, Y. Xia, Langmuir 2001, 17, 6344.

Das könnte Ihnen auch gefallen