Вы находитесь на странице: 1из 5

Materials Chemistry and Physics 129 (2011) 501–505

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Preparation of Ag-doped TiO2 nanoparticles by a miniemulsion


method and their photoactivity in visible light illuminations
Youji Li ∗ , Mingyuan Ma, Wei Chen, Leiyong Li, Mengxiong Zen
College of Chemistry and Chemical Engineering, Jishou University, Jishou 46000, Hunan, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Visible-light-driven photocatalysts based on silver-doped TiO2 (Ag–TiO2 ) nanoparticles were successfully
Received 26 June 2009 synthesized by a miniemulsion method and characterized by X-ray diffraction (XRD), scanning elec-
Received in revised form 9 March 2011 tron microscope (SEM), X-ray photoelectron spectroscopy (XPS), BET surface area analysis and UV–vis
Accepted 22 April 2011
diffuse reflectance spectroscopy (DRS). Degradation of methylene blue (MB) was applied to evaluate
photocatalytic activity of samples. The results show that Ag doping showed a controlling effect on the
Keywords:
transformation of titania from anatase to rutile. A red shift occurred in the absorption edge of titania with
Nanostructures
the certain Ag-doped amount. Moreover, the addition of Ag resulted in a higher Brunauer–Emmett–Teller
Chemical synthesis
Electron microscopy
(BET) surface area as well as a larger average pore size of TiO2 nanoparticles. The specific surface area
Oxidation increased with the Ag-doped amount to reach a maximum (86.3 m2 g−1 ) at Ag/Ti molar ratio of 0.8%
and then decreased with further increase of the Ag-doped amount. The Ag–TiO2 nanoparticles could
effectively photodegrade MB under visible light irradiation and the obtained maximum reaction con-
stant (kapp = 0.007 min−1 ) was three times higher than that of pure TiO2 (kapp = 0.002 min−1 ) when the
Ag-doped amount was 0.8%. The commendable visible photoactivities of Ag–TiO2 photocatalysts are pre-
dominantly attributable to simultaneous effects of Ag deposits by the acting as electron traps, enhancing
the MB adsorption on the Ag–TiO2 surface, occurring red shift of the absorption edge and decreasing band
gap.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction Ag+ ions makes a dramatic improvement in the performance of TiO2


photocatalysts, however, the aforementioned silver based photo-
Recently, TiO2 has been widely studied for its wide application catalysts function only under UV light. There are only few articles
in photocatalysis [1], solar cells [2], hydrogen production [3] and reported about the photoactivity of Ag–TiO2 in visible light illumi-
catalyst support [4,5]. However, its wide band gap (anatase: 3.2 eV; nations. Additionally, researchers discovered that materials in the
rutile: 3.0 eV), the extensive exploitation of TiO2 created an expec- nano-size domain exhibit a unique surface chemical reactivity for
tation to use merely 3–4% UV light of the whole radiant solar energy. the destructive absorption of acid gas and chemical warfare agents
Extensive efforts have been made in the development of TiO2 pho- [27,28]. Moreover, this unique surface reactivity of nanomaterials
tocatalysts that can efficiently utilize solar or indoor light to extend was attributed to the high surface areas and the presence of defects,
the activating spectrum from UV to visible light [6–10]. These stud- edges, and corners, which are available of improving photoactiv-
ies include doping metal ions into the TiO2 lattice [11,12], dye ity of TiO2 in visible light illuminations. There are different routes
photosensitization on the TiO2 surface [1,13–16], and deposition of that can be used to synthesize titanium dioxide. These include
noble metals [17–21]. In particular, noble metal-modified semicon- flame synthesis [29], chemical vapor deposition [30], precipitation
ductor nanoparticles become of current importance for maximizing [31] and the alkoxide sol–gel method [32]. The miniemulsifica-
the efficiency of photocatalytic reactions. In recent years, there tion technique, having a controlled reaction rate, is a promising
have been some reports on the preparation of Ag+ -doped TiO2 films method to synthesize nanomaterials with high surface area [33].
and nanoparticles that degrade some textile dyes (methyl orange, Therefore, it is rationally more desirable to synthesize visible-light-
crystal violet, and methyl red) in aqueous medium [22–24]. Fur- driven photocatalysts based on silver-doped TiO2 nanoparticles by
thermore, silver halides, such as AgBr/SiO2 and AgCl catalysts, have a miniemulsion method.
also been used in photocatalysis [25,26]. It appears that doping with In this paper, silver-doped titania nanoparticles were prepared
by a miniemulsion method with the hydrolysis of Ti(OBun)4 in the
presence of silver nitrate. The photoreactivity of the synthesized
∗ Corresponding author. Tel.: +86 13762157748. materials was evaluated by degradation of methylene blue, cho-
E-mail address: bcclyj@163.com (Y. Li). sen as a standard model molecule, was performed under visible

0254-0584/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2011.04.055
502 Y. Li et al. / Materials Chemistry and Physics 129 (2011) 501–505

light irradiation. The samples were characterized using DRS, XRD,


nitrogen adsorption–desorption and SEM measurements.

2. Experimental

2.1. Preparation of Ag–TiO2 nanoparticles

The tetrabutyltitanate, cyclohexanol, hexadecyl trimethyl ammonium bromide,


triethylamine and acetone were used as titanium source, impregnant, surfactant,
precipitator, and lotion, respectively. Oil–water components system was con-
firmed in the experiments of preparing Ag–TiO2 nanoparticles by a miniemulsion
method. The particularly synthesizing process was followed: tetrabutyltitanate
(10 ml) was used to produce titanium dioxide particles by hydrolysis and miniemul-
sion dispersed in cyclohexanol (50 ml) was added to a premixed solution among, Fig. 1. Experimental apparatus for photocatalytic reaction.
cyclohexanol (50 ml), a certain weight of hexadecyl trimethyl ammonium bro-
mide, distilled water and silver nitrate at room temperature. This solution had been
aged for 12 h to obtain silver-doped TiO2 particles with using triethylamine as pre- 3. Results and discussion
cipitator. These particles were recovered and washed three times with acetone.
Centrifugation was used to recover the particles. The particles were then allowed to 3.1. Catalyst characterization
dry in an oven at 80 ◦ C for 2 h, followed by calcinations at different temperature for
2 h. Four kinds of Ag–TiO2 were prepared by varying the molar ratio of silver relative
to titania after vigorous stirring for several hours. The samples were denominated
Liquid nitrogen absorption-desorption experimental results are
X% Ag–TiO2 (X% denoted the molar ratio of silver to titanium). presented in Table 1, the Brunauer–Emmett–Teller (BET) surface
area was strongly dependent on Ag-doped amount. Obviously, the
specific surface areas of the doped TiO2 samples are higher than
2.2. Catalyst characterization
that of the undoped TiO2 . When the Ag-doped amount was 0, 0.5,
The crystalline phase was identified by X-ray diffraction (Bruker, Germany) 0.8, 1.0 and 2.0%, the SBET of Ag–TiO2 was 47.7, 72.5, 86.3, 69.5
using Cu-K␣ as radiation and were analyzed from 20◦ to 80◦ (2). Nitrogen and 60.2 m2 g−1 , respectively. It can be seen that SBET increased
adsorption–desorption isotherms were recorded at 77 K (Micromeritics ASAP 2020)
with the Ag-doped amount to reach a maximum at 0.8% and then
and the specific surface areas were determined by the Brunauer–Emmett–Teller
(BET) method. The diffuse reflectance UV–Vis spectra of the samples were deter-
decreased with further increase of the Ag-doped amount. It is
mined by a Cray-300 UV/VIS spectrophotometer at room temperature. BaSO4 was understandable that the higher the Ag-doped amount, the lower
used as the standard in all measurements. The morphology of the prepared samples the BET surface area because of sintering of the smaller parti-
were observed by SEM (JSM-5600LV, Japan). Samples were prepared by ultrasonic cles into the bigger particles. That is why TiO2 catalysts at 2.0% of
dispersion of the powder in alcohol prior to testing at the room temperature. The
Ag-doped amount have the least surface area (60.2 m2 g−1 ). Mean-
average particle size of TiO2 was measured by photon correlation spectroscopy (PCS)
with an argon ion laser (Lexel Laser Inc. Model 95-2) operated at 514.4 nm and while, the Ag-doped amount has a profound effect on the pore
200 mW. The chemical states of silver and titanium were studied by XPS measure- size distribution of samples calcined at 500 ◦ C as shown in Fig. 3.
ments, using a VG Scientific ESCALAB Mark spectrometer (England), where the X-ray The narrow pore size distribution of pores is in range of 3–12 nm,
photoelectron spectra were reference to the C1s peak (Eb = 285.0 eV). and the average pore size of Ag–TiO2 increased with an increase
in Ag-doped amount. As presented in Table 1, doped samples still
2.3. Photoactivity measurements retained intra-particles mesopores, while the intra-particles meso-
pores almost decreased for the undoped samples. Fig. 4 illustrates
Methylene blue (MB) was chosen as a model organic compound to evaluate the
the X-ray diffraction (XRD) profiles of silver-doped TiO2 samples
photoactivity of the prepared Ag–TiO2 nanoparticles. Fig. 1 shows a scheme of the
experimental apparatus. The main component of the system was the reactor, on calcined progressively at 400, 500, and 600 ◦ C for 2 h and undoped
which there was a 1000 W high-pressure mercury lamp to provide the irradiance TiO2 calcined at 550 ◦ C. It can be seen that there were sharp and
source with internal stirring by a magnetic stirrer. There is a filter to assure visible strong peaks of anatase phase at 2 = 25.3◦ in XRD patterns of all
light in photocatalytic experiments at a distance of 20 cm from the top. The initial
samples. There appeared no peaks for rutile (2 = 27.5◦ ) phases in
concentration of methylene blue in stirred aqueous solution was fixed at 2 mg l−1
with a catalyst loading of 5 g l−1 . To determine the change of methylene blue concen-
all 0.8% Ag–TiO2 samples. It indicated that the phase transforma-
tration in solution during the photocatalytic process, a few milliliters of the solution tion from anatase to rutile did not take place when 0.8% Ag–TiO2
was taken from the reaction mixture, subsequently centrifuged, filtered through a powder was calcined at temperatures below 600 ◦ C. It meant that
Millipore filter (pore size 0.22 ␮m) to separate the catalysts, and loaded in a UV–vis silver could control the phase transformation from anatase to rutile.
spectrometer (JascoV-500, Japan). For accurate evaluation of photocatalytic degra-
It was well known that the occurrence of the anatase-to-rutile tran-
dation rate of MB, each experiment was repeated at least twice. The absorption peak
at 660 nm was adopted to make calibration curve as shown in small illustration of sition actually depends on several factors (e.g. preparation method,
Fig. 2. The methylene blue concentration was calculated by using a calibration curve particle sizes, heating rate, doping level). Similar findings were also
according to the absorbance intensity at 660 nm in photocatalytic process. reported by other authors [18,20]. They had found that silver pre-

30
a b
Concentration (mg·L–1)

20

10 Y2 = 0.997
y = 6.1705x

0
500 550 600 650 700 0 2 4 6

Wavelength (nm) Absorption intensity (a.u.)

Fig. 2. UV–vis spectrum of methylene blue (a) and relationship between concentration of MB and its absorption intensity of UV–vis at 660 nm (b).
Y. Li et al. / Materials Chemistry and Physics 129 (2011) 501–505 503

Table 1
The characteristics of titanium dioxide with Ag doping.

Samples Surface areas (m2 g−1 ) Mean pore size (nm) Band energy (eV) Kinetic constant (mine−1 )

TiO2 47.4 – 3.16 0.002


0.5% Ag–TiO2 72.5 3 3.08 0.005
0.8% Ag–TiO2 86.3 5 3.00 0.007
1.0% Ag–TiO2 69.5 8 2.98 0.006
2.0% Ag–TiO2 60.2 12 2.92 0.004

als with photon energies. Therefore, it is absolutely imperative to


e e: 2%Ag- TiO 2 include this technique to examine for the photoactivity of synthe-
b c d d: 1%Ag- TiO 2 sized photocatalysts. Fig. 5 demonstrates the DRS spectra profiles
c: 0.8%Ag- TiO 2 of titanium dioxide with various amount of silver calcined at 500 ◦ C
for 2 h. Here, we perceived a noticeable shift of the optical absorp-
b: 0.5%Ag- TiO 2
tion edges of the doped TiO2 systems toward the visible regions of
dV(dR) / (cc/g)

a: TiO 2
the solar spectrum. Notably, this redshift toward the longer wave-
lengths originates from the band gap narrowing of titanium dioxide
by silver doping and the band gap energy of the doped samples
was determined from the equation, Eg = 1239.8/, where  is the
wavelength (nm) of the exciting light [36]. The band gap energies
of the samples were listed in Table 1. Because the doped samples
have lower band gap energies than the undoped TiO2 (3.16 eV),
these photocatalysts are, therefore, likely to operate under visi-
a
ble light illumination. Additionally, it is evidently observed that
1 2 4 8 16 32 64 the Ag-doped amount can result the profound effect on UV–vis
Diameter size(nm) absorption spectra of the Ag–TiO2 catalysts. The absorption of the
sample in the visible range was of 400–600 nm and increased with
Fig. 3. The pore size distribution curve of TiO2 calcined at 500 ◦ C for 2 h. the increase of the Ag-doped amount in TiO2 . In the SEM micro-
graphs shown in Fig. 6, only titanium dioxide particles can be seen
although the doped Ag is coloured for the presence of the silver.
vented the phase transformation as well as had an inhibition effect
Thus it can be concluded that either the amount of silver parti-
on the growth of anatase crystallite. However, the opposite results
cles is too low, so that the probability to find a silver particle in
about promoting effect of Ag doping on the phase transformation
SEM is small, or that the size of the silver particles is too small, so
were appeared in some articles [34,35]. Additionally, no notice-
that the resolution of SEM is not sufficient to detect them. The par-
able peaks of silver oxide and silver metal were observed in the
ticle size of the Ag–TiO2 nanoparticles was measured by PCS. The
X-ray diffractograms perhaps due to the microscale of Ag-doped
TiO2 particles prepared without the addition of silver (Fig. 6a) were
amount. Diffuse reflectance UV–vis spectroscopy directly provides
aggregated and the average size of the particles was 30 nm. When
some insight into the interactions of the photocatalyst materi-
Ag-doped amount was 0.8%, the Ag–TiO2 particles were spheri-
cal and the aggregation was reduced (Fig. 6b). The particles were
homogeneously dispersed and the size was about 20–30 nm with
the average of 25 nm. However, the excess silver (2.0%) deterio-
rated the dispersion of the particles and the aggregation appeared

Fig. 5. Diffuse reflectance UV–vis spectra of TiO2 samples, prepared by a miniemul-


Fig. 4. XRD patterns of titania powders calcined at different temperatures. sion method at 500 ◦ C for 2 h.
504 Y. Li et al. / Materials Chemistry and Physics 129 (2011) 501–505

Fig. 6. Scanning electron micrographs of Ag–TiO2 calcined at 500 ◦ C for 2 h. (a) TiO2 . (b) 0.8% Ag–TiO2 . (c) 2% Ag–TiO2 .

Ag+ Ag 0 a b
Ti3+ Ti4+

Intensity (arb.units)
Intensity (arb.units)

Ag3d5/2 Ag3d3/2

Ti2p3/2

Ti2p1/2

360 365 370 375 380 456 460 464 468


Binding energy (eV) Binding energy (eV)

Fig. 7. High-resolution XPS spectra of the 0.8% Ag–TiO2 nanoparticles: Ag 3d (a) and Ti 2p (b).

again (Fig. 6c). The size of the TiO2 particles was about 30–40 nm trates the profound effect of Ag-doped amount on the methylene
with the average of 35 nm. X-ray photoelectron spectroscopy (XPS) blue degradation on Ag–TiO2 in visible light illuminations. After
experiments were performed to elucidate both the chemical state 450 min photodegradation reaction excitated by visible light irradi-
of the silver particles and the titania structure. As shown in Fig. 7a, ation, 82–100% methylene blue were decomposed on the Ag–TiO2
in terms of the chemical bonding states of Ag2 O (Ag+ , 367.8 eV) and samples, while only 24% methylene blue were degraded on the
Ag (Ag0 , 368.2 eV) in the Ag 3d XPS peaks, respectively [37,38], the pure TiO2 prepared without the addition of Ag. The degradation of
presence of both Ag0 and Ag+ species was determined by measuring methylene blue on pure TiO2 catalyst under visible light irradiation
the XPS of the best Ag–TiO2 catalyst. In general, advisable amount was attributed to the photosensitization process [40]. Anatase tita-
of Ag doping is available of enhancing the catalytic property of the nia could not be directly excitated by visible light due to its 3.2 eV
material because the presence of both Ag0 and Ag+ species facilitate band gap, but the photobleaching reaction still occurred under
the charge separation (electron–hole) and suppress the recombina- TiO2 /dye/visiblelight system due to dye molecule acting as photo-
tion of the photoexcited charge carriers. Additionally, Fig. 7b shows
the narrow scans for Ti 2p peaks located at 464.8 eV (Ti 2p1/2) and
458.8 eV (Ti 2p3/2) of the Ag–TiO2 photocatalysts which represent 100
the existence of two peaks at 457.5 eV and 458.9 eV matching the
trivalent and tetravalent states of Ti, respectively. Acted as active
Residue rate of methylene blue (%)

center, vacancy sites (Ti3+ sites) can trap the photogenerated elec- 80
trons in the conduction band and prevent the recombination of
electron–hole pairs under UV. So these Ti3+ sites play important TiO2
roles in the photooxidation of organic compound on the TiO2 pho- 2%Ag- TiO2
60
tocatalysts [39]. 0.5%Ag- TiO2
1%Ag- TiO2
0.8%Ag- TiO2
3.2. Photocatalytic activity 40

Prior to photocatalytic reactions, the suspension was mag-


netically stirred in the dark for 60 min to establish an 20
adsorption–desorption equilibrium at room temperature. Photoac-
tivities for methylene blue in the dark in the presence of the
photocatalysts and under visible light irradiation in the absence 0
of the photocatalysts were also evaluated. It was found that there 0 100 200 300 400 500
was no degradation for the methylene blue in the dark and in the
Time (mine)
presence of the photocatalyst. Also no degradation was observed
for methylene blue when the solution was placed under visible light Fig. 8. Decomposition of methylene blue in aqueous solution by TiO2 process in
radiation without the addition of photocatalyst powder. Fig. 8 illus- visible light illuminations with various Ag incorporated content.
Y. Li et al. / Materials Chemistry and Physics 129 (2011) 501–505 505

sensitizer. The electron from the excited dye molecule was injected simultaneous effects of Ag deposits by the acting as electron traps,
into the conduction band (CB) of the TiO2 , and the cation radical enhancing the MB adsorption on the Ag–TiO2 surface, occurring red
formed at the dye surface quickly undergoes degradation reaction shift of the absorption edge resulting from a decrease in band gap.
to oxidise methylene blue into products. Additionally, it greatly
affected the Ag–TiO2 photoactivity that the addition concentration References
of silver was varied from 0.5 to 2.0%. The optimal photoactivity of
the catalyst was achieved for 0.8% of Ag+ ion. This means that 0.8% [1] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. A: Chem. C: Pho-
tochem. Rev. 1 (2000) 1.
Ag+ ion concentration effectively suppresses the recombination of [2] B. O’Regan, M. Gratzel, Nature 353 (1991) 737, 10.
the photogenerated charge-carriers on the surface of the catalyst. [3] A. Fujishima, K. Honda, Nature 238 (1972) 37.
Unfortunately, an increase in Ag+ ion concentration from 0.8 to 2.0% [4] S. Koutsopoulos, S.B. Rasmussen, K.M. Eriksen, R. Fehrmann, Appl. Catal. A: Gen.
306 (2006) 142.
has a deleterious effect on the photoactivity of the catalysts. Con- [5] J.T. Miller, A.J. Kropf, Y. Zha, J.R. Regalbuto, L. Delannoy, C. Louis, E. Bus, J.A. van
ceivably, this happens because of the creation of recombination Bokhoven, J. Catal. 240 (2) (2006) 222.
centers of charge-carriers at a higher loading of dopant concentra- [6] R. Asahi, T. Morikawa, T. Ohwaki, Science 293 (2001) 269.
[7] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Science 297 (2002) 2243.
tion [11]. Meanwhile, 0.8% Ag+ ion concentration can result in high
[8] M. Anpo, M. Takeuchi, J. Catal. 216 (2003) 505.
surface area of catalyst so that a large number of reactant molecules [9] T. Ihara, M. Miyoshi, Y. Iriyama, O. Matsumoto, S. Sugihara, Appl. Catal. B:
are absorbed and undergo subsequent oxidation reactions. The Environ. 42 (2003) 403.
decomposition behavior of methylene blue in aqueous solution by [10] Y.B. Xie, C.W. Yuan, Appl. Catal. B 46 (2003) 251.
[11] S.D. Mo, L.B. Lin, J. Phys. Chem. Solids 55 (1994) 1309.
doped TiO2 process was adequately described by pseudo first-order [12] W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669.
reaction kinetics and the calculated rate constants are depicted [13] K. Vinodgopal, D.E. Wynkoop, P.V. Kamat, Environ. Sci. Technol. 30 (1996) 1660.
in Table 1. The photodecomposition rate constant of methylene [14] J. Zhao, T. Wu, K.Wu.K. Oikawa, H. Hidaka, N. Serpone, Environ. Sci. Technol. 32
(1998) 2394.
blue by doped TiO2 process enhanced from 0.002 to 0.007 min−1 [15] T. Wu, G. Liu, J. Zhao, H. Hidaka, N. Serpone, J. Phys. Chem. B 102 (1998) 5845.
with an increase in the silver incorporated content up to a certain [16] G. Liu, T. Wu, T. Lin, J. Zhao, Environ. Sci. Technol. 33 (1999) 1379.
level (optimum silver incorporated content) and then decreased. [17] A. Sclafani, J.-M. Herrmann, J. Photochem. Photobiol. A 113 (1998) 181.
[18] B.H. Dambar, J.K. Amal, Kenneth, J. Colloid Interface Sci. 311 (2007) 514.
The rate of organic oxidation is dependent on the concentration [19] V. Subramanian, E. Wolf, P. Kamat, J. Phys. Chem. B 105 (2001) 11439.
of organic molecules and oxidising agents in the suspension [2,4]. [20] J.Y. Park, J.J. Yun, C.H. Hwang, I.H. Lee, Mater. Lett. 64 (24) (2010) 2692.
The excess incorporated content of silver-laden particles may cover [21] V. Vamathevan, R. Amal, D. Beydoun, G. Low, S. McEvoy, J. Photochem. Photo-
biol. A 148 (2002) 233.
active sites on the TiO2 surface, and thereby reducing photodegra- [22] P. Falaras, I.M. Arabatzis, T. Stergiopoulos, M.C. Bernard, Int. J. Photoenergy 5
dation efficiency. (2003) 123.
[23] K. Gupta, A. Pal, C. Sahoo, Dyes Pigments 69 (2006) 224.
[24] Y. Liu, C.Y. Liu, Q.H. Rong, Z. Zhang, Appl. Surf. Sci. 220 (2003) 7.
4. Conclusions [25] N. Kakuta, N. Goto, H. Ohkita, T. Mizushima, J. Phys. Chem. B 103 (9) (1999)
5917.
[26] A. Currao, V.R. Reddy, M.K. van Veen, R.E.I. Schropp, G. Calzferri, Photochem.
Ag–TiO2 nanoparticles were prepared by a miniemulsion Photobiol. Sci. 3 (2004) 1017.
method using Ti(OBun)4 and Ag(NO3 ) as starting materials. Ag dop- [27] C.L. Carnes, K.J. Klabunde, Chem. Mater. 14 (4) (2002) 1807.
ing could control the transformation of titania from anatase to [28] G.M. Medine, V. Zaikovskii, K.J. Klabunde, J. Mater. Chem. 14 (2004) 757.
[29] S.E. Pratsinis, J. Aerosol Sci. 27 (1996) 153.
rutile. With increasing of the Ag-doped amount, the absorption [30] Z. Ding, X. Hu, P.L. Yue, G.Q. Lu, P.F. Greenfield, Catal. Today 68 (2001) 178.
edge of Ag–TiO2 shifted to longer wavelength, and the band gap [31] H.D. Nam, B.H. Lee, S. Kim, C. Jung, J. Lee, S. Park, J. Appl. Phys. 37 (1998) 4606.
energy of Ag–TiO2 decreased. With a suitable amount (0.5–2.0%), [32] K.Y. Jung, S.B. Park, J. Photochem. Photobiol.: A Chem. 127 (1999) 119.
[33] K.C. Song, J.H. Kim, Powder Technol. 107 (2000) 270.
the Ag dopant increases the specific surface area of TiO2 pow- [34] H.E. Chao, Y.U. Yun, H.U. Xingfang, A. Larbot, J. Eur. Ceram. Soc. 23 (9) (2003)
der. In photodegradation MB solution, the photocatalytic activity of 1457.
Ag–TiO2 increased with the increase of Ag-doped molar ratio from [35] X.B. Li, L.L. Wang, X.H. Lu, J. Hazard. Mater. 177 (1–3) (2010) 639.
[36] A.R. Gandhe, J.B. Fernandes, Bull. Catal. Soc. India 4 (2005) 131.
0 to 0.8%, but declined with the further increase to 2.0%. The optimal
[37] X.Y. Gao, S.Y. Wang, J. Li, Y.X. Zheng, R.J. Zhang, P. Zhou, Y.M. Yang, L.Y. Chen,
molar ratio for photocatalytic activity of Ag–TiO2 was found as 0.8% Thin Solid. Films 455/456 (2004) 438.
with the high BET surface area of 86.3 m2 g−1 , of which rate con- [38] F.X. Bocka, T.M. Christensenb, S.B. Riversc, L.D. Doucettea, R.J. Lad, Thin Solid
stant (kapp = 0.007 min−1 ) was three times higher than that of pure Films 468 (2004) 57.
[39] H. Liu, H.T. Ma, X.Z. Li, W.Z. Li, M. Wu, X.H. Bao, Chemosphere 50 (2003) 39.
TiO2 (kapp = 0.002 min−1 ). The commendable visible photoactivi- [40] M.S.S. Hyung, R.C. Jae, J.H. Hoe, M.K. Sang, C.B. Young, J. Photochem. Photobiol.
ties of Ag–TiO2 photocatalysts are predominantly attributable to A: Chem. 163 (2004) 37, 12.

Вам также может понравиться