Está en la página 1de 6

DOI: 10.1002/cssc.

200900100

Titania Nanocrystals and Adsorptive Nanoporous Polymer


Composites: An Enrichment and Degradation System
Yonglai Zhang,[a] Shu Wei,[a] Wei Zhang,[b] Yi-Jun Xu,[b] Sen Liu,[a] Dang Sheng Su,*[b] and
Feng-Shou Xiao*[a]

Composites of titania nanocrystals and adsorptive nanoporous the nanoporous polymers. The composites exhibit nanoporous
polymers are synthesized via a two-step solvothermal pathway. structures, high surface areas, and excellent stabilities. Experi-
These composites are developed as bifunctional materials for mental results show that the composites have excellent ad-
both the adsorption of organic compounds and photocatalysis, sorptive properties for organic compounds in water and air
providing an enrichment and degradation system for the re- and high photocatalytic activities under UV irradiation. These
moval of environmental pollutants. The anatase-phase titania bifunctional materials may have promising applications in envi-
nanoparticles are homogeneously dispersed in the matrix of ronmental protection.

Introduction

In recent years, protection of the environment has become a Herein, we report the first in situ synthesis of titania nano-
significant issue in sustainable chemistry. Therefore, considera- crystals and nanoporous polydivinylbenzene (PDVB) compo-
ble efforts have been devoted towards investigating various sites. These composites are efficient bifunctional materials for
adsorbents for treating toxic pollutants in air and water. Gener- the adsorption of organic compounds and photocatalytic deg-
ally, nanoporous polymer and carbon materials can act as effi- radation.
cient absorbents and could be used to solve environmental
problems.[1–7] Porous polymer materials show high adsorptive
capacities for organic compounds owing to their oleophilic
frameworks.[1] They differ from inorganic porous materials,
Results and Discussion
which can be regenerated by simple calcinations; porous poly- Synthesis
mer and carbon materials can be reused only after using a
Titania nanocrystals and nanoporous polymer composites were
proper method to desorb the adsorbed pollutants, which
synthesized via a two-step solvothermal pathway (Figure 1).
strongly impedes their industrial applications. An ideal adsorb-
Here, solvothermal polymerization was considered a facile
ent would decompose pollutants and self-regenerate for recy-
method for the preparation of nanoporous polydivinylbenzene
cling use. On the other hand, photocatalytic degradation of or-
(PDVB) containing a loose network.[35] Divinylbenzene mono-
ganic pollutants is also a fascinating solution to environmental
mers were polymerized in the presence of THF and tetrabutyl
problems.[8–15] However, the low frequency of collision between
titanate (TBT) in the first solvothermal treatment. After the in-
photocatalysts and organic pollutants lowers the reaction rate
troduction of a desired amount of water, TBT was hydrolyzed,
of mineralization. Furthermore, secondary pollution caused by
forming amorphous titania nanoparticles in the confined space
the byproducts produced during the photocatalytic process
of the polymeric skeleton. After a second solvothermal treat-
also limits the practical applications of photocatalysts.
ment, nanocrystals of titania were formed and homogeneously
A promising way to overcome the disadvantages of both ad-
dispersed in the system. Finally, titania nanocrystals and nano-
sorbents and photocatalysts is to combine their respective ad-
vantages in a controlled fashion. Some pioneering works, for [a] Y. Zhang, S. Wei, S. Liu, Prof. F.-S. Xiao
example, the loading of TiO2 particles onto various adsorbents College of Chemistry and
(activated carbon, zeolite, Al2O3, mesoporous silicate, and natu- State Key Laboratory of Inorganic Synthesis and Preparative Chemistry
Jilin University
ral clay),[16–23] the synthesis of mesoporous titania,[24–29] the fab-
Qianjin Street 2699, Changchun 130012 (PR China)
rication of mesoporous TiO2/SiO2 and TiO2/C composites,[30–34] Fax: (+ 86) 431 85168624
and the synthesis of porous polymers as photocatalysts,[5] have E-mail: fsxiao@mail.jlu.edu.cn
been reported. The as-obtained materials revealed improved [b] Dr. W. Zhang, Dr. Y.-J. Xu, Dr. D. S. Su
properties for both the adsorption and photodegradation of Department of Inorganic Chemistry
Fritz Haber Institute of the Max Planck Society
organic pollutants. An evaluation of the adsorptive properties
Faradayweg 4–6, 14195 Berlin (Germany)
of these bifunctional materials, however, is still lacking. Thus, it E-mail: dangsheng@fhi-berlin.mpg.de
is a challenge to develop highly efficient adsorptive materials Supporting information for this article is available on the WWW under
containing excellent photocatalytic activities. http://dx.doi.org/10.1002/cssc.200900100.

ChemSusChem 2009, 2, 867 – 872  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 867
D. S. Su, F.-S. Xiao et al.

Table 1. Physicochemical properties of titania nanoparticles and nanopo-


rous polymer composites prepared under different conditions.

Sample TiO2 content [wt %] SBET Vp dp


TGA[a] Mass reduction[b] ACHTUNGRE[m2 g 1] ACHTUNGRE[cm3 g 1] [nm]
P-T-0.25 – 5.7 651 0.88 9.3
P-T-0.5 – 9.8 617 0.89 9.2
P-T-1 16.4 17.2 543 1.03 11.4
P-T-2 30.6 31.0 500 0.94 11.8
P-T-3 45.2 45.8 424 1.04 22.1
P-T-4 – 47.8 354 1.13 25.4
P-T-1[c] – – 484 1.02 11.8

[a] Results calculated from thermogravimetric analysis (TGA). [b] Results


Figure 1. Synthesis of titania nanocrystals and nanoporous polymer compo- calculated from mass reduction after calcination at 650 8C for 3 h. [c] Sam-
sites. TBT: tetrabutyl titanate; (P)DVB: (poly)divinylbenzene ples exposed to UV irradiation (125 W) for 7 d. SBET: Brunauer–Emmett–
Teller surface area; Vp : total pore volume; dp : average pore diameter.

porous polymer composites were obtained after evaporation


of the solvents (THF, water, and n-butanol, generated from the
hydrolysis of TBT). The samples are labeled as P-T-x, where x
stands for the mass ratio of TBT to DVB in the starting solu-
tions.

Characterization of the samples


Nitrogen adsorption–desorption isotherms and pore size distri-
bution curves of the P-T-x samples (x = 0.25–4) are shown in
Figure 2. The P-T-0.25 sample showed a typical type IV iso-
therm with a hysteresis loop (P/P0 at 0.7–0.9, Figure 2 a, curve
A), indicating a meso-structured composite. Correspondingly,
the pore size, calculated from the desorption curve, is 9.3 nm.
Composites with different TiO2 content also showed nitrogen Figure 3. Wide-angle XRD patterns of (A) P-T-0.25, (B) P-T-0.5, (C) P-T-1, (D) P-
uptakes (P/P0 at 0.7–1.0), giving surface areas of 354– T-2, (E) P-T-3, and (F) P-T-4 (solid squares: anatase; solid triangle: brookite).
651 m2 g 1 and pore sizes of 9.2–25.4 nm (Table 1).
Figure 3 shows the wide-angle X-ray diffraction (XRD) pat-
terns of the P-T-x samples, showing the predominant anatase solid squares). A small amount of brookite phase (2q = 30.88,
phase (2q = 25.38, 38.08, 48.08, 55.08, 62.68, 68.78, and 75.08, solid triangle) was present in all samples. The broad peak be-
tween 108 and 308 corresponds to the amorphous polymeric
framework.
The morphology and porosity of the composites were char-
acterized by scanning electron microscopy (SEM). As typical ex-
amples, SEM images and energy-dispersive X-ray (EDX) spectra
of P-T-1 and P-T-3 are shown in Figure 4. The P-T-1 composite
showed a monolithic morphology (inset of Figure 4 a). The
high-magnification image shows that P-T-1 is highly porous,
with a pore size in the mesopore range (Figure 4 b). The P-T-3
composite shows characteristics similar to P-T-1 except for the
presence of large cavities in the macropore range (Figure 4 d
and e). The EDX spectra (Figure 4 c and f) only shows the pres-
ence of carbon, oxygen, and titanium in the samples. Quantita-
tive results revealed that the atomic ratio of oxygen to titani-
Figure 2. a) N2 adsorption–desorption isotherms and b) pore size distribution um is close to the ideal value of 2. In addition, the two sam-
curves of (A) P-T-0.25, (B) P-T-0.5, (C) P-T-1, (D) P-T-2, (E) P-T-3, and (F) P-T-4. ples showed different Ti-Ka,b and O–K content (Figure 4 c and
The isotherms of B–F have been offset by 300, 600, 900, 1200, and f), in agreement with the titania content shown in Table 1.
1500 cm3 g 1, respectively, along the vertical axis for clarity. The pore size dis-
Figure 5 shows the microstructure investigations for P-T-3.
tribution curves of B–F have been offset by 1.5, 3, 4.5, 6, and 7.5 cm3 g 1, re-
spectively, and each of them was estimated by the Barrett–Joyner–Halenda Titania nanoparticles with an average diameter of 10 nm uni-
model from a desorption branch of the isotherm. formly dispersed in the PDVB framework. Only carbon, titani-

868 www.chemsuschem.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2009, 2, 867 – 872
Titania Nanocrystals and Adsorptive Nanoporous Polymer Composites

The scanning transmission elec-


tron microscopy (STEM) images
in Figure 5 c show that titanium,
oxygen, and carbon are homo-
geneously distributed in the
composite materials.
Figure 6 shows the UV/vis
spectra of P-T-1 and P-T-3 as well
as the commercially available
photocatalyst P25 for compari-
son. A significant increase in the
absorption at wavelengths
below 410 nm can be assigned
to the intrinsic band gap absorp-
tion of anatase and rutile TiO2
for P25. The absorption edge
shifts towards shorter wave-
lengths for P-T-1 and P-T-3 sam-
ples (380 nm) owing to the band
gap absorption of anatase and
brookite TiO2 (Figure 3).[36] The
slight absorption band for P-T-1
between 380 and 500 nm can
be assigned to the absorption of
PDVB.

Stabilities of the samples


The thermal stability of the com-
posites was investigated by ther-
mogravimetric analysis (TGA).
Figure 7 a shows a significant
Figure 4. a,b) SEM images, and c) EDX spectra of P-T-1; d,e) SEM images, and f) EDX spectra of P-T-3. mass loss starting at 370 8C for
P-T-1. This is much higher than
pure PDVB (ca. 270 8C), which
um, and oxygen were detected (see the electron energy loss
spectroscopy (EELS) spectrum, inset of Figure 5a), indicating
the presence of polymers and titania, consistent with the EDX
spectroscopy results. Figure 5 b shows highly crystallized titania
nanoparticles. It is known that anatase titania (space group =
I41/amd) has lattice parameters of a = b = 3.78 , c = 9.51 ,
and a = b = g = 908 (see inset of Figure 5b). In the high-resolu-
tion transmission electron microscopy (HRTEM) image, the lat-
tices of the (101) and (011) planes of anatase titania were iden-
tified and their interplanar spacings are 0.35 nm. The diameter
of the particles is ca. 10 nm, as indicated by the circled area.
Between the particles is amorphous PDVB. Similar observations
were made for the P-T-x samples with various titania content.
As shown in the Supporting Information, titania nanoparticles
with diameters of 5–10 nm homogeneously distributed in the
P-T-0.5 and P-T-1 samples (see HRTEM images). It is worth
noting that the size of the titania nanoparticles does not line-
arly increase in proportion to the titania content. No large tita- Figure 5. a) Transmission electron microscopy (TEM), and b) high-resolution
(HR)TEM images of P-T-3; the inset in (a) is the EELS spectrum; the inset in
nia particles were found even though the TiO2 content was
(b) is the crystallography model of anatase titania where Ti atoms are in
very high (ca. 45 % for P-T-3, Table 1) because the PDVB net- white and O atoms are in gray; c) single (S)TEM image, Ti, O, and C elemen-
work hindered the agglomeration of the titania nanoparticles. tal maps of the sample (from left to right).

ChemSusChem 2009, 2, 867 – 872  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 869
D. S. Su, F.-S. Xiao et al.

Figure 6. UV/vis spectra of P-T-1, P-T-3, and P25.

was prepared as a comparable sample, indicating that the in-


troduction of titania into the PDVB framework greatly im-
proves the thermal stability of the composites. In addition, the
P-T-1 sample gave a large mass loss (83.6 %), indicating a tita-
nia content of 16.4 %. When the titania content was increased,
mass loss for the P-T-2 and P-T-3 samples started at 385 and
410 8C, respectively, further confirming the effect of titania on
the thermal stability of the composites. The P-T-2 and P-T-3
samples gave relatively small mass losses (69.4 and 54.8 %, re-
spectively) owing to the high titania content. The titania con-
tent calculated from the TGA curves is similar to that calculat-
ed from mass loss after calcination at 650 8C for 3 h in air
(Table 1). The stability of these samples under UV irradiation
Figure 7. a) TGA curves of P-T-1, P-T-2, P-T-3, and pure PDVB. b) N2 isotherms
was also evaluated by exposing the P-T-1 composite to UV irra- and pore size distribution curves of P-T-1 before and after exposure to UV ir-
diation (125 W) for 7 days. As shown in Figure 7 b , the P-T-1 radiation (125 W) for 7 d. Isotherm and pore size distribution curves have
sample shows similar nitrogen adsorption isotherms and pore been offset by 200 cm3 g 1 and 1.5 cm3 g 1, respectively, along the vertical
axis for clarity.
size distribution curves before and after UV irradiation for
7 days, indicating the excellent stability of the porous struc-
ture. The FTIR spectra (see Supporting Information) further
confirmed the high stability. For example, after irradiation for 7 750 m2 g 1, respectively) showed relatively low adsorptive rates
days, the IR spectrum of P-T-1 did not change. compared with the P-T-3 sample (424 m2 g 1). This phenomen-
on is possibly owing to the different hydrophobicities of the
samples. In the second step, the mixture was exposed to UV ir-
Adsorptive properties and photocatalytic activities
radiation for a further 180 min. Interestingly, an abrupt de-
The adsorptive properties and photocatalytic activities of the crease in the concentration of the RhB solutions was observed
composites were evaluated by a two-step procedure and com- after the saturated adsorption for the P-T-1 and P-T-3 samples
pared with P25 and Amberlite XAD-4 adsorptive resin (Fig- under UV irradiation, which is assigned to the photocatalytic
ure 8 a). Rhodamine-B (RhB) was chosen as a probing pollutant degradation of the RhB adsorbed by the composite materials
for both adsorption and photocatalytic experiments. Firstly, an and the re-adsorption of the RhB in the solution. In addition,
adsorptive procedure was performed in the dark for 180 min. the decrease in concentration for P-T-3 (45.8 % titania) is more
Both P-T-1 and P-T-3 samples showed similarly high adsorptive obvious than that of P-T-1 (17.2 % titania), owing to the rich-
capacity and adsorptive rate. More than 68 % of RhB in the ness of titania in the P-T-3 sample. The P25 sample also gave
aqueous solution was adsorbed by the P-T-1 sample (65 % for an abrupt decrease in concentration owing to its photocatalyt-
P-T-3). Following adsorption for 180 min, the adsorptive capaci- ic decomposition. On the contrary, the XAD-4 resin did not
ty reaches saturation. On the contrary, P25 showed negligible show an abrupt decrease in concentration because of the ab-
adsorption owing to the low specific surface area (ca. sence of photocatalytic activity. The P-T-1 and P-T-3 samples
45 m2 g 1). The XAD-4 resin showed a relatively low adsorptive showed superior adsorptive properties and good photocatalyt-
rate and nearly 25 % of RhB was adsorbed after 180 min. The ic activities compared to the P25 and XAD-4 samples. The P25
P-T-1 and XAD-4 samples with large BET surface areas (543 and sample only exhibited photocatalytic activity without any ad-

870 www.chemsuschem.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2009, 2, 867 – 872
Titania Nanocrystals and Adsorptive Nanoporous Polymer Composites

matrix, which lacks titania nanoparticles. Particularly, the aque-


ous solution remained clear during UV irradiation, indicating
no desorption of RhB.
The gas phase adsorptive properties were tested using ace-
tone and water as probing molecules (Figure 9). Interestingly,
all of the composites showed large adsorptive capacities for

Figure 9. Adsorptive capacity of acetone and water as probing molecules in


gas phase over P-T-x (x = 0.25–4) and XAD-4.

acetone (1.24–0.60 g g 1). Compared with the adsorptive resin


XAD-4, the P-T-0.25 and P-T-0.5 samples gave higher adsorp-
tive capacities, indicating superior adsorptive properties. A fur-
ther increase in the titania content (P-T-x, x = 1–4) resulted in
the decrease of adsorptive capacity for acetone, which is at-
Figure 8. a) Concentration change of RhB solutions under adsorptive (black tributed to the decrease in polymer content and BET surface
line) and photocatalytic (blue line) processes over P-T-1, P-T-3, P25, and
area. On the contrary, the adsorptive capacity for water in-
XAD-4 resin, respectively. b) Concentration change of RhB solutions under
adsorptive and photocatalytic process over P-T-3 sample. (Content of RhB creased with the titania content owing to the decrease in hy-
adsorbed in P-T-3 during photocatalytic process was evaluated by extracting drophobicity. Generally, the organic framework and the porous
RhB from the hybrid of RhB and P-T-3 in ethanol. Inset: absorption change structure are favorable for the formation of hydrophobic surfa-
of extracted RhB in ethanol).
ces.[37, 38] As a result, a larger organic polymer content contrib-
utes to a higher hydrophobicity. The XAD-4 resin also shows
significant hydrophobicity and this feature might explain the
sorptive abilities; whereas XAD-4 exhibited general adsorptive low adsorptive rate for the XAD-4 resin. The high adsorptive
properties without photocatalytic activity. Thus, our compo- capacity for organic compounds in the gas phase indicates the
sites combining adsorptive and photocatalytic properties are potential use of these composite materials for the enrichment
much more efficient for treating pollutions in aqueous solu- and degradation of volatile organic compounds.
tions than single photocatalysts or adsorbents.
To identify the photocatalytic activity for the RhB adsorbed
in the composite materials after full adsorption, the hybrids of
P-T-3 and RhB with different photodegradation times were ex-
Conclusions
tracted into ethanol, respectively. The concentration change of
RhB in ethanol represents the content change of the adsor- Titania nanocrystals and nanoporous adsorptive polymer com-
bate. As shown in Figure 8 b, the RhB solution (20 ppm) was posites have been prepared by a two-step solvothermal path-
fully adsorbed by the P-T-3 sample after 60 min in the dark. way. These composites were developed as bifunctional materi-
The mixture was then exposed to UV irradiation and the pho- als for both adsorption and photocatalysis, and designed as an
tocatalytic degradation of the RhB adsorbed in the P-T-3 enrichment and degradation system for the elimination of en-
sample occurred. The concentration of RhB extracted into the vironmental pollutants. Experimental results show that these
ethanol showed a significant decrease owing to the photocata- composites exhibit excellent adsorptive properties and high
lytic decomposition of RhB. Most of the RhB decomposed fol- photocatalytic activities. The bifunctional materials could po-
lowing UV irradiation for 180 min. The slight residue was de- tentially be used in the decontamination of organic pollutants
rived from the adsorbed RhB molecules in the polymeric from water and the atmosphere.

ChemSusChem 2009, 2, 867 – 872  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 871
D. S. Su, F.-S. Xiao et al.

Experimental Section [3] a) Y. Meng, D. Gu, F. Zhang, Y. Shi, H. Yang, Z. Li, C. Yu, B. Tu, D. Zhao,
Angew. Chem. 2005, 117, 7215; Angew. Chem. Int. Ed. 2005, 44, 7053;
DVB was purchased from Tianjin-Guangfu Chemical Corp., and THF, b) F. Zhang, Y. Meng, D. Gu, Y. Yan, C. Yu, B. Tu, D. Zhao, J. Am. Chem.
azobisisobutyronitrile (AIBN), TBT, and acetone were purchased Soc. 2005, 127, 13508.
from Beijing Chemical Corp. Amberlite XAD-4 resin was purchased [4] a) F. Svec, J. M. J. Frechet, Science 1996, 273, 205; b) E. C. Peters, F. Svec,
J. M. J. Frechet, Adv. Mater. 1999, 11, 1169.
from Aldrich. P25 was purchased from Degussa Co. All chemicals
[5] a) P. Kuhn, K. Kruger, A. Thomas, M. Antonietti, Chem. Commun. 2008,
were used as-received.
5815; b) X. Wang, K. Maeda, X. Chen, K. Takanabe, K. Domen, Y. Huo, X.
In a typical synthesis of titania nanocrystals and adsorptive nano- Fu, M. Antonietti, J. Am. Chem. Soc. 2009, 131, 1680.
porous polymer composites, the monomer of DVB (2 g) was dis- [6] C. Liang, Z. Li, S. Dai, Angew. Chem. 2008, 120, 3754; Angew. Chem. Int.
solved in THF (10 mL) containing TBT (y g, y = 0.5–8), and the initia- Ed. 2008, 47, 3696.
tor AIBN (0.05 g) was added. After stirring for 4 h at RT, the trans- [7] J. Lee, S. Han, T. Hyeon, J. Mater. Chem. 2004, 14, 478.
parent solution was transferred into an autoclave. After solvother- [8] M. R. Hoffmann, S. T. Martin, W. Choi, D. W. Bahnemann, Chem. Rev.
mal treatment at 100 8C for 24 h, deionized water (10 mL) was 1995, 95, 69.
added into the autoclave to hydrolyze the TBT. After another solvo- [9] A. Maldotti, A. Molinari, R. Amadelli, Chem. Rev. 2002, 102, 3811.
[10] M. I. Litter, Appl. Catal. B 1999, 23, 89.
thermal treatment at 100 8C for 24 h and 180 8C for 24 h, a solid
[11] O. M. Alfano, D. Bahnemann, A. E. Cassano, R. Dillert, R. Goslich, Catal.
monolith was obtained following slow evaporation of the solvents
Today 2000, 58, 199.
at RT. The samples were labeled as P-T-x, where x stands for the [12] S. Malato, J. Blanco, A. Vidal, C. Richter, Appl. Catal. B 2002, 37, 1.
mass ratio of TBT to DVB in the starting solutions. A PDVB sample [13] G. Palmisano, V. Augugliaro, M. Pagliaro, L. Palmisano, Chem. Commun.
without addition of titanium was prepared in the same conditions 2007, 3425.
as the P-T-x samples. [14] W. Zhao, W. Ma, C. Chen, J. Zhao, Z. Shuai, J. Am. Chem. Soc. 2004, 126,
Adsorption and photocatalytic experiments were performed by a 4782.
two-step procedure. Typically, the sample (50 mg) was added to an [15] S. Han, S. H. Choi, S. S. Kim, M. Cho, B. Jang, D. Y. Kim, J. Yoon, T. Hyeon,
aqueous solution (50 mL) containing RhB (20 ppm or 100 ppm) Small 2005, 1, 812.
under vigorous stirring. Firstly, the mixture was placed in the dark [16] N. Takeda, T. Torimoto, S. Sampath, S. Kuwabata, H. Yoneyama, J. Phys.
Chem. 1995, 99, 9986.
for 1 or 3 h to evaluate the adsorptive properties. Then, the solu-
[17] T. Torimoto, S. Ito, S. Kuwabata, H. Yoneyama, Environ. Sci. Technol.
tion was irradiated by a 125 W mercury lamp for another 3 h to 1996, 30, 1275.
evaluate the photocatalytic activities. The concentration of RhB [18] B. Tryba, A. W. Morawski, M. Inagaki, Appl. Catal. B 2003, 41, 427.
was determined at different intervals by UV-vis absorption. [19] J. Matos, J. Laine, J. M. Herrmann, Appl. Catal. B 1998, 18, 281.
Nitrogen adsorption isotherms were measured with a Micromeritics [20] Y. Yu, J. C. Yu, C. Y. Chan, Y. K. Che, J. C. Zhao, L. Ding, W. K. Ge, P. K.
Tristar 3000 analyzer at 77 K. Samples were degassed for 10 h at Wong, Appl. Catal. B 2005, 61, 1.
150 8C before the measurements were recorded. Powder XRD data [21] L. Zou, Y. Luo, M. Hooper, E. Hu, Chem. Eng. Process. 2006, 45, 959.
were collected on a Rigaku D/MAX 2550 diffractometer with Cu Ka [22] S. Fukahori, H. Ichiura, T. Kitaoka, H. Tanaka, Appl. Catal. B 2003, 46, 453.
radiation (l = 1.5418 ). SEM experiments were performed on a Hi- [23] S. Zhan, D. Chen, E. Jiao, C. Tao, J. Phys. Chem. B 2006, 110, 11199.
[24] S. Zhan, D. Chen, X. Bazso, V. Zollmer, A. Richardt, I. Dekany, Chemo-
tachi S4800 (2.0 kV) equipped with EDX detectors. TEM experi-
sphere 2008, 70, 538.
ments were performed on a Philips CM 200 LaB6 (200 kV) and a [25] H. Shibata, T. Ogura, T. Mukai, T. Ohkubo, H. Sakai, M. Abe, J. Am. Chem.
Philips CM200 FEG (200 kV) equipped with EELS. TG analysis was Soc. 2005, 127, 16396.
carried out on a NETZSCH STA 449C with a heating rate of [26] H. Li, Z. Bian, J. Zhu, Y. Huo, H. Li, Y. Lu, J. Am. Chem. Soc. 2007, 129,
20 8C min 1 from RT to 800 8C. The concentrations of the RhB solu- 4538.
tions were measured by a Shimadzu (UV-2450) UV/vis spectropho- [27] E. Stathatos, T. Petrova, P. Lianos, Langmuir 2001, 17, 5025.
tometer. Measurements of adsorptive capacity for acetone and [28] Z. Y. Yuan, T. Z. Ren, B. L. Su, Adv. Mater. 2003, 15, 1462.
water were estimated from the change in sample weight before [29] G. Liu, Y. Zhao, C. Sun, F. Li, G. Q. Lu, H. M. Cheng, Angew. Chem. 2008,
and after exposure of probing vapor to the samples under vacuum 120, 4592; Angew. Chem. Int. Ed. 2008, 47, 4516.
[30] a) A. Nakajima, K. Hashimoto, T. Watanabe, Langmuir 2000, 16, 7044;
conditions (ca. 1 torr) at RT. Before adsorption, the samples were
b) S. Zhan, D. Chen, X. Jiao, Y. Song, Chem. Commun. 2007, 2043.
evacuated at 120 8C for 3 h to remove adsorbates. FTIR spectra [31] S. Perathoner, P. Lanzafame, R. Passalacqua, G. Centi, R. Schlogl, D. S. Su,
were performed on a IFS 66 V/S (Bruker) IR spectrometer in the Microporous Mesoporous Mater. 2006, 90, 347.
range 400–4000 cm 1, and the samples were measured until the [32] R. Liu, Y. Ren, Y. Shi, F. Zhang, L. Zhang, B. Tu, D. Zhao, Chem. Mater.
final pressure was 4  10 3 mbar. 2008, 20, 1140.
[33] P. D. Southon, R. F. Howe, Chem. Mater. 2002, 14, 4209.
[34] a) F. X. Llabrs i Xamena, P. Calza, C. Lamberti, C. Prestipino, A. Damin, S.
Acknowledgements Bordiga, E. Pelizzetti, A. Zecchina, J. Am. Chem. Soc. 2003, 125, 2264;
b) S. Usseglio, P. Calza, A. Damin, C. Minero, S. Bordiga, C. Lamberti, E.
Pelizzetti, A. Zecchina, Chem. Mater. 2006, 18, 3412.
This work was supported by the State Basic Research Project of
[35] a) Y. Zhang, S. Wei, F. Liu, Y. Du, S. Liu, Y. Ji, T. Yokoi, T. Tatsumi, F. S. Xiao,
China (2009CB623507), the National Natural Science Foundation Nano Today 2009, 4, 135; b) Y. Zhang, S. Wei, H. Zhang, S. Liu, F. Nawaz,
of China (20773049), and the PPP. We gratefully acknowledge Dr. F. S. Xiao, J. Colloid. Interf. Sci. 2009, DOI: 10.1016/j.jcis.2009.07.050.
Jian Zhang at the Fritz-Haber-Institute of the Max Planck Society- [36] S. D. Mo, W. Y. Ching, Phys. Rev. B 1995, 51, 13023.
[37] L. Jiang, Y. Zhao, J. Zhai, Angew. Chem. 2004, 116, 4438; Angew. Chem.
for helpful discussions.
Int. Ed. 2004, 43, 4338.
[38] X. Zhang, F. Shi, J. Niu, Y. G. Jiang, Z. Q. Wang, J. Mater. Chem. 2008, 18,
621.
Keywords: adsorption · composites · nanoporous materials ·
photochemistry · titanium

[1] D. C. Kennedy, Environ. Sci. Technol. 1973, 7, 138. Received: April 22, 2009
[2] J. A. Leenheer, Environ. Sci. Technol. 1981, 15, 578. Published online on August 31, 2009

872 www.chemsuschem.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2009, 2, 867 – 872

También podría gustarte