You are on page 1of 9

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3

Available at www.sciencedirect.com

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

Silver-modified mesoporous TiO2 photocatalyst for water


purification

Zhigang Xiong a, Jizhen Ma a, Wun Jern Ng b, T. David Waite c, X.S. Zhao a,*
a
Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, Singapore 117576,
Singapore
b
Nanyang Environment & Water Research Institute, Nanyang Technological University, Nanyang Avenue, Singapore 639798, Singapore
c
School of Civil and Environmental Engineering, The University of New South Wales, Sydney, NSW 2052, Australia

article info abstract

Article history: Mesoporous anatase (TiO2) was modified with silver (Ag) nanoparticles using a photore-
Received 22 July 2010 duction method. Performance of the resulting TiO2eAg nanocomposites for water purifi-
Received in revised form cation was evaluated using degradation of Rhodamine B (RhB) and disinfection of
5 December 2010 Escherichia coli (E. coli) under ultraviolet (UV) irradiation. The composites with different Ag
Accepted 20 December 2010 loadings were characterized using physical adsorption of nitrogen, X-ray diffraction, X-ray
Available online 24 December 2010 photoelectron spectroscopy and UVeVisible diffuse reflectance spectroscopic techniques.
The results showed that metallic Ag nanoparticles were firmly immobilized on the TiO2
Keywords: surface, which improved electron-hole separation by forming the Schottky barrier at the
Silver nanoparticles TiO2eAg interface. Photocatalytic degradation of RhB and inactivation of E. coli effectively
Mesoporous titanium dioxide occurred in an analogical trend. The deposited Ag slightly decreased adsorption of target
Photocatalysis pollutants, but greatly increased adsorption of molecular oxygen with the latter enhancing
Disinfection production of reactive oxygen species (ROSs) with concomitant increase in contaminant
photodegradation. The optimal Ag loadings for RhB degradation and E. coli disinfection
were 0.25 wt% and 2.0 wt%, respectively. The composite photocatalysts were stable and
could be used repeatedly under UV irradiation.
ª 2011 Elsevier Ltd. All rights reserved.

1. Introduction ultraviolet (UV) irradiation. This has often led to incomplete


mineralization of organic pollutants (Tachikawa et al., 2007).
The titanium dioxide (TiO2) photocatalyst is gaining growing To effectively improve the mineralization process, a means is
research interest (Chen and Mao, 2007; Han et al., 2009; needed to enhance adsorption of organic pollutants or dis-
Hoffmann et al., 1995; Linsebigler et al., 1995; Xu et al., 2007; solved O2 onto the surface of TiO2 (Kang et al., 2008; Mogyorosi
Xu and Langford, 1995; Yu et al., 2003). Upon UV light irradia- et al., 2002). However, due to its poor affinity towards organic
tion, TiO2 generates highly reactive oxygen species (ROSs) and pollutants and the low surface area of TiO2 nanoparticles, the
valence holes, which can degrade most recalcitrant pollutants, adsorption of organic pollutants on TiO2 surface is relatively
including non-biodegradable dyes (Han et al., 2009), phenols low, resulting in slow photocatalytic degradation rates. To
and their derivatives (Xu et al., 2007), and inactivate various overcome this shortcoming, TiO2 nanoparticles have been
microorganisms, such as bacteria, viruses, and tumor cells supported on various porous solids (Hidaka et al., 1990; Kang
(Sunada et al., 2003b). However, the lifetime of ROSs, such as et al., 2008; Mogyorosi et al., 2002; Xu and Langford, 1995).
O  
2 , HO , HOO and H2O2, is relatively short, particularly under Using advanced porous materials for water treatments has

* Corresponding author. Tel.: þ65 65164727; fax: þ65 67791936.


E-mail address: chezxs@nus.edu.sg (X.S. Zhao).
0043-1354/$ e see front matter ª 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2010.12.019
2096 w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3

been demonstrated to be a very promising approach to water biodegradable dye, Rhodamine B (RhB), and a gram-negative
purification (Pan et al., 2010; Zhao et al., 2010). The materials bacterium, E. coli, were used as probes to investigate the
can be synthesized via template (Tian et al., 2002; Yu et al., photocatalytic detoxification and disinfection properties of
2003) and solegel methods (Pinna and Niederberger, 2008; the composite materials under UV irradiation. The effect of Ag
Raveendran et al., 2008). Mesoporous TiO2 photocatalysts loadings on catalyst performance and the role of Ag nano-
have been observed to display an improved performance in the particles in the photocatalytic degradations were investigated.
degradation of chlorophenols and dyes under UV irradiation
(Beyers et al., 2005; Schattka et al., 2002). The porous structure
of mesoporous TiO2 may also provide an adequate environ- 2. Materials and methods
ment for the adsorption of dissolved O2 on the oxygen vacan-
cies (Epling et al., 1998). It is recognized that reduction of the 2.1. Materials
adsorbed O2 is a rate-limiting step in the photocatalytic
degradation of organic pollutants (Hoffmann et al., 1995; Titanium isopropoxide, silver nitrate, RhB, ethanol, and
Sadeghi et al., 1996), and so increase in O2 adsorption on TiO2 phosphate buffer saline of analytical grade were used without
surfaces would enhance charge separation, and so generate further purification. Pluronic surfactant poly(ethylene glycol)-
more ROSs. block-poly(propylene glycol)-block-poly(ethylene glycol) (EO20
Another factor that greatly restricts the photocatalytic PO70EO20, P123) was purchased from Aldrich. Deionized water
activities of TiO2 is the low quantum yield of excitons due to the was used in this work.
fast electron-hole (eehþ) recombination. Methods to effec-
tively improve the photocatalytic activity of TiO2 have been 2.2. Preparation and modification of mesoporous TiO2
widely explored over the past decade. Direct doping of an
element, optimization of morphology and facets, and adding The synthesis of mesoporous anatase TiO2 was similar to that
additional scavengers are amongst the most effective reported previously (Yu et al., 2003). 6.0 g of P123 and 9.0 mL of
approaches (Chen and Mao, 2007; Hoffmann et al., 1995; Xu titanium isopropoxide were dissolved in 60 mL of ethanol at
et al., 2007; Xu and Langford, 1995). In addition, deposition ambient temperature. After the suspension was stirred for 1 h,
of metal nanoparticles with a large work function, such as 30 mL of H2O was added and stirred for another 1 h. After
Ag (Tran et al., 2006), Pt (Kowalska et al., 2008), and Au (Bannat complete evaporation of ethanol and water, white solids were
et al., 2009; Subramanian et al., 2001; Wang et al., 2008), onto obtained. The solids were calcined at 350  C for 5 h in air with
TiO2 surface has been found to efficiently retard the eehþ a heating rate of 0.5  C/min. The resultant product was
recombination because of the Schottky barrier formed at the denoted as m-TiO2.
metalesemiconductor interface. Here, the metal nanoparticles Modification of mesoporous TiO2 with Ag nanoparticles
act as a mediator in storing and shuttling photogenerated was conducted using a photoreduction method (Tran et al.,
electrons from the TiO2 surface to an acceptor. Iliev et al. (Iliev 2006). 1.0 g of m-TiO2 was suspended in 300 mL of water, fol-
et al., 2006) demonstrated that deposition of Pt or Ag on the lowed by sonication for 10 min to completely disperse the solid.
surface of TiO2 greatly enhanced the photocatalytic degrada- Then, AgNO3 solution was added to the suspension. After
tion of oxalic acid due to the increased separation of eehþ and stirring for 20 min in the dark, the mixture was irradiated with
higher rate of O2 reduction. Similarly, Li et al. (2007) homoge- a 125 W high-pressure mercury lamp (Philips, Belgium) for
nously embedded gold particles onto the TiO2 framework, 4.0 h. The resulting pale-gray solids were collected, washed
which significantly improved photocatalytic degradation of with water and dried at ambient temperature. No Ag ion was
phenol and chromium in solution. Other organic pollutants, like detected in the filtrate indicating that all Ag had been
methanol (Ismail et al., 2009), methylene blue (Wang et al., 2008) completely loaded onto the m-TiO2. Different volumes of
and methyl orange (Arabatzis et al., 2003), could also be more AgNO3 solution with a concentration of 6.27  102 M were
efficiently degraded by the metaleTiO2 nanocomposite. Reac- used to vary the loading of Ag. The samples thus obtained are
tive oxygen species, and particularly hydroxyl radicals (HO∙), denoted as m-TiO2eAgx%, where x% indicates the weight
produced by the irradiated metaleTiO2 has been considered to percentage of Ag on the composite particles.
be the dominant species contributing to the degradation of
various organic pollutants (Arabatzis et al., 2003). 2.3. Characterization
The metal-modified TiO2 could also be used for the inacti-
vation of microorganisms. Sunada et al. (Sunada et al., 2003a) X-ray diffraction (XRD) patterns were collected on a Shimadzu
compared the photocatalytic disinfection of Escherichia coli in XRD-6000 X-ray diffractometer using Cu Ka irradiation oper-
pure TiO2 and Cu-loaded TiO2 with the latter displaying an ated at 40 kV and 30 mA. The average crystallite size was
enhanced bactericidal activity. Silver modification of TiO2 has estimated using the Scherrer equation based on the (101)
been shown to enhance the bactericidal activity of UV-irradi- reflection of anatase. Adsorptionedesorption isotherms of N2
ated TiO2 by about seventy folds as a result of both improved were measured on a Micromeritics ASAP2020 system at 77 K.
microorganism adsorption to the particle surface and suppres- The BrunauereEmmetteTeller surface area (SBET) was calcu-
sion of eehþ recombination (Zhang et al., 2010). The formed lated in the relative pressure range of P/P0 ¼ 0.05  0.2. The
ROSs was found to lead to perturbation of various cellular pore volume was estimated at P/P0 ¼ 0.97. Diffuse-reflectance
processes and eventually to bacterial death (Chen et al., 2009b). spectra were recorded on a Shimadzu 3100 UVeVis-NIR
In the present study, Ag-modified mesoporous anatase spectrometer equipped with an integrating sphere ISR-3100
TiO2 was synthesized via a photoreduction method. A non- using BaSO4 as the reference. High-resolution transformation
w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3 2097

electron microscope (HRTEM) images were obtained on a Per-


kineElmer JEM-2010F operated at 200 kV. X-ray photoelectron
spectra (XPS) were collected on an AXIS HIS (Kratos Analytical
Ltd., UK) instrument with an Al Ka irradiation source
(1486.7 eV) operated at 15 kV and 10 mA. Photoluminescence
spectra were collected at room temperature on a PTI lumi-
nescence spectrometer with a solid accessory.

2.4. Photocatalytic degradation of RhB

Photocatalytic degradation of RhB was carried out in a semi-


batch swirl flow monolithic-type photoreactor (Zhou et al.,
2006). Before UV irradiation, a suspension containing 300 mL
of 8.6  103 mM RhB solution and 0.15 g of solid photocatalyst
was sonicated for 5 min and stirred for 20 min in the dark to
allow sorption equilibrium. The mixture was then irradiated
with a 125 W high-pressure mercury lamp. At a given time
interval of irradiation, 5 mL aliquots were withdrawn and
filtered through a 0.45 mm membrane. The concentration of
RhB in the filtrate was analyzed using a Shimadzu 1601 PC
UVeVis spectrophotometer.

2.5. Photocatalytic disinfection of E. coli

The antibacterial experiments were carried out in a sterilized


100 mL glass conical flask containing 50 mL 7.7  108 CFU/mL E.
coli and 0.5 g/L solid photocatalyst under room temperature.
The suspension was irradiated by two 8 W Black light Blue (BLB)
lamps (200 mW/cm2). After a given time interval, 50 mL of the
suspension was sampled, and diluted with sterile phosphate Fig. 1 e (A) Nitrogen adsorption isotherms and (B) pore size
buffer saline (PBS, 0.03 mol/L, pH 7.2) to yield appropriate distribution curves of mesoporous TiO2 modified with Ag
starting concentrations (ensuring cells were well suspended by nanoparticles of loadings of (a) 0%, (b) 0.1%, (c) 0.25%, (d)
vortex mixing before transfer). The suspensions were then 0.5%, (e) 1.0%, (f) 2.0%, and (g) 3.0%.
spread on Petri Dishes which had been previously coated with
nutrient agar. After being incubated at 37  C for 18 h, the formed
colonies were counted and used to calculate the amount of
active cells. Each data point was determined three times and displayed a type IV isotherm with an H3 hysteresis loop, indi-
was used as an average. Control experiments without the cating that mesoporous materials are involved (Sing et al., 1985).
photocatalyst were also conducted for comparison. The relatively narrow pore-size distribution curves indicate
a single pore-size system (Fig. 1B). Table 1 summarizes the pore
2.6. Stability of photocatalyst parameters of the samples. The BET surface area of m-TiO2 was
calculated to be about 166 m2/g, higher than that reported
Stability of the photocatalyst was evaluated as follows: 0.15 g previously (Yu et al., 2003). Modification with Ag gradually
of the photocatalyst m-TiO2eAg 0.25% was suspended in decreased the surface area with increasing loading of Ag - from
300 mL of 8.0  103 mM RhB solution. After a given time 0.1% to 2.0%. Pore volume also decreased - from 0.43 to
interval of UV irradiation, 5 mL aliquots were withdrawn, 0.33 cm3/g on increasing Ag loading from 0.1% to 2.0%. The
filtered and analyzed on the UVeVis spectrophotometer. After above results indicate that Ag had been loaded onto the m-TiO2
RhB was completely degraded, the suspension was further sample. As the amount of silver increased, more Ag particles
irradiated for 20 min. Then, 15 mL of stock solution (0.16 mM) formed on the surface of mesoporous TiO2. The Ag particles
of RhB was supplied to restore the initial concentration of RhB covered some sorption sites and lowered the pore volume of the
(8.0  103 mM). The suspension was equilibrated again for TiO2. However, they did not remarkably affect the pore struc-
20 min in the dark before again being irradiated with UV light. ture (Fig. 1B) (Liu et al., 2008). For sample m-TiO2eAg 3.0%, an
obvious decrease in pore size can be seen suggesting that the Ag
particles blocked some of the pores.
3. Results and discussion Fig. 2 shows the XRD patterns of the samples. All resolved
peaks can be indexed as the (101), (004), (200) and (211) reflec-
3.1. Characterization tion of pure anatase (JCPDS No. 21-1272) (Chen et al., 2009a; Yu
et al., 2003). The crystal sizes were calculated from the (101)
Fig. 1A shows the adsorptionedesorption isotherms of TiO2 and plane and are given in Table 1. It can be seen that within the Ag
Ag-deposited TiO2 photocatalysts. It is seen that all samples loading range of 0.1%e3.0%, the crystal size of the mesoporous
2098 w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3

Table 1 e Physical parameters of TiO2 and Ag-deposited


TiO2.
Crystallite SBET Pore Average
size (nm)a (m2/g) volume pore
(cm3/g)b size (nm)c

m-TiO2 6.7 166 0.43 11.3


m-TiO2eAg 0.1% 7.1 162 0.38 10.8
m-TiO2eAg 0.25% 6.7 157 0.38 11.5
m-TiO2eAg 0.5% 6.9 149 0.38 11.5
m-TiO2eAg 1.0% 6.9 139 0.36 11.6
m-TiO2eAg 2.0% 6.6 128 0.33 11.0
m-TiO2eAg 3.0% 6.6 125 0.29 10.1

a Calculated using Scherrer equation.


b Pore volume at P/P0 ¼ 0.97.
c Estimated from the desorption branches using the BJH model.

anatase was not altered. No reflection peak owing to metallic


Ag can be seen in Fig. 2, presumably due to the small Ag particle Fig. 3 e UVeVisible diffuse reflectance spectra of
size. The TEM image (Fig. S1) confirmed that the size of the Ag mesoporous TiO2 modified with Ag nanoparticles of
particles in sample m-TiO2eAg 3.0% with the highest Ag loadings ranging from 0 (bottom) to 3.0% (top). The label on
loading was smaller than 1.5 nm. The size of the Ag particles in top of each curve is the wavelength of the highest
other samples was too small to be distinguished from the TiO2 absorption.
background using the TEM technique. The Ag 3d XPS spectra of
Ag-modified mesoporous TiO2 samples (Fig. S2) showed two
increased with Ag loading, indicating the deposition of Ag
peaks, at 368 and 374 eV, assigned to Ag 3d5/2 and 3d3/2,
nanoparticles on the TiO2 surface (Linsebigler et al., 1995).
respectively (Guin et al., 2007). No peak due to AgD is seen.
However, the maximum wavelength of the surface plasmon
Considering that no Ag ion was detected in the filtrate after the
absorption varied from one sample to another. As Ag loading
photoreduction process, it thus can be concluded that all AgD
increased from 0.1% to 3.0%, the wavelength of the highest
ions added into the solution had been completely reduced to
surface plasmon absorption gradually shifted from 468 nm to
metallic Ag nanoparticles.
498 nm, accompanied with the broadening of the absorption
The UVeVis diffuse reflectance spectra of all samples are
peaks. It is known that surface plasmon resonance (SPR)
shown in Fig. 3. The m-TiO2 samples exhibited strong
strongly depends on the size and uniformity of the nano-
absorption in the UV region with a sharp absorption edge
particles (He et al., 2002) with the SPR peak becoming broader
located at about 400 nm (Zhang et al., 2009). Upon loading
and shifting to longer wavelengths as the particle size
different amounts of Ag, no change in the absorption edge is
increases, and/or when the particle size distribution becomes
seen. A typical surface plasmon absorption of Ag nano-
broad. The results in Fig. 3 suggest that as Ag loading gradually
particles at around 400e550 nm gradually emerged and
increased, more Ag nanoparticles aggregated on the TiO2
surface. Such an aggregation resulted in a decrease in the pore
volume as evidenced by the data in Table 1.

3.2. Photocatalytic detoxification and disinfection under


UV light irradiation

The photocatalytic properties of the mesoporous anatase TiO2


and Ag-deposited TiO2 photocatalysts were evaluated by
examination of RhB dye degradation and Gram-negative
bacterium E. coli inactivation under UV irradiation. Before
commencement of light irradiation, the mixtures of RhB or
E. coli solution and solid photocatalyst were stirred for 20 min
in the dark to achieve complete sorption equilibrium.
As shown in Fig. 4A, minor removal of RhB occurred with
catalyst m-TiO2eAg in the absence of UV irradiation. This
removal was due to adsorption of the dye on the solid particle
surface. In the absence of a photocatalyst but under UV irra-
diation, however, a slow degradation was observed. In the
Fig. 2 e XRD patterns of mesoporous TiO2 before and after presence of m-TiO2, the concentration of RhB in suspension
modification with Ag nanoparticles of different loadings. decreased rapidly with irradiation time. The data could be
The standard XRD pattern of anatase is also included for satisfactorily described by the first-order kinetic equation, ln
reference. (C/C0) ¼ kappt, where kapp is the apparent rate constant, C and
w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3 2099

et al., 2009; Sharma et al., 2009). When the suspension was


irradiated with UV in the presence of m-TiO2eAg 3%, the
inactivation of E. coli was greatly enhanced. Analyzing using
first-order kinetics (Chen et al., 2009c), the total rate constants
decreased in the order m-TiO2eAg 3% (0.102 min1) > m-TiO2
(0.055 min1) > m-TiO2eAg 3% (dark, 0.040 min1) > UV
(0.024 min1) > m-TiO2 (dark, 0.0073 min1). The contribution of
Ag inactivation to the total rate constant could be estimated to
be about 0.033 min1.After eliminating the Ag inactivation
effect, the photocatalytic disinfection effect can be assessed as
exhibiting a net rate constant of k ¼ 0.062 min1, larger than
that of m-TiO2. This trend is similar to that observed for the
photocatalytic degradation of RhB. The time-course of inacti-
vation of E. coli however is substantially more complicated than
the degradation of RhB, possibly as a result of the combined
effects of the antimicrobial activity of Ag itself and the photo-
catalytic disinfection effect of the TiO2eAg composite with the
synergic effect of the two processes leading to a significant
improvement of antibacterial activity (Liu et al., 2008). Silver is
known for its antimicrobial properties with some investigators
proposing that Ag nanoparticles may attach to the surface of
the cell membrane and so disturb permeability and respiration
functions of the cell (Kvitek et al., 2008). Notwithstanding this,
the bactericidal mechanism of Ag nanoparticles remains to be
fully understood (Sharma et al., 2009). The photocatalytic
disinfection process has also been widely investigated. Chen
and coworkers (Chen et al., 2009c) compared the photocatalytic
degradation of formaldehyde and inactivation of E. coli with
TiO2 and concluded that both processes depended on the ROS
production rate. This is also apparent in the present system
since the generation of ROSs is an important phenomenon in
Fig. 4 e Photocatalytic degradation of (A) RhB and (B) E. coli the TiO2eAg suspension. Disinfection performance of the
in different system. The initial concentration of RhB and photocatalyst can be evaluated by the analogy method based
E. coli were 8.6 3 10L3 mM and 7.7 3 108 CFU/mL, on dye degradation under UV irradiation since the latter is
respectively. easier to measure quantitatively.

3.3. Effect of Ag loading

C0 are the concentrations of RhB at time t ¼ t and t ¼ 0, Deposition of Ag nanoparticles on TiO2 surface not only
respectively. The rate constant of the photocatalytic degrada- enhances photocatalytic reactions (Sung-Suh et al., 2004) but
tion with m-TiO2 was calculated to be about 0.044 min1. When
m-TiO2eAg 0.25% was used, the RhB degradation was
enhanced exhibiting a rate constant of 0.069 min1, compa-
rable to 0.073 min1 of P25. The absorption spectra of RhB were
gradually decreased with irradiation time with no shift of the
main absorption peak, suggesting complete cleavage of RhB
chromophores. Although the UV energy is also adequate for
RhB excitation, the photosensitized degradation, as evidenced
by a concomitant wavelength shift of the absorption band to
shorter wavelengths (Sung-Suh et al., 2004; Wu et al., 1998), can
be neglected compared to the photocatalytic degradation.
The antibacterial properties of the catalysts were studied
using E. coli as a model strain. In the presence of UV irradiation
and without a photocatalyst (Fig. 4B) or in the presence of
m-TiO2 without UV light, the number of viable E. coli cells
decreased slowly. With m-TiO2eAg 3% and without UV irradi-
ation, the viable E. coli number decreased sharply in the first Fig. 5 e Effect of Ag loading on the photocatalytic
60 min, then became stable. This behavior is different to RhB degradation of RhB (C) and E. coli (B), and on the
degradation and may be partially attributed to the intrinsic photoluminescent intensity ratio between Ag-deposited
disinfection properties of the deposited Ag nanoparticles (Lv TiO2 and TiO2 (¤).
2100 w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3

photocatalytic degradation of RhB (about 0.25 wt%) was much


larger than that for maximal eehþ separation. The results
suggest that the deposition of Ag nanoparticles on TiO2 might
have other roles other than to simply depress eehþ
recombination.
Considering the deposited Ag particles covered the surface
of TiO2, the adsorption affinity of the solid catalyst towards both
dissolved oxygen and RhB would most likely be altered. As
shown in Fig. 6, upon Ag deposition, RhB adsorption decreased
with increasing Ag loading. Since RhB molecules mainly
adsorbed on TiO2 surface with the diethylamino group (Wu
et al., 1998) or the carboxyl group (Zhao et al., 2008), deposi-
tion of Ag particles on TiO2 would inevitably covered some
sorption sites, resulting in a decrease of the RhB adsorption
Fig. 6 e Effect of Ag loading on the adsorption of RhB and (Fig. 7). Control experiments (Fig. S4A) showed that the
the surface area using O2 as an adsorptive. The initial adsorption of RhB on m-TiO2eAg 0.25% surface followed
concentration of RhB was 8.6 3 10L3 mM. a Langmuir model with the maximum absorption capacity at
about 5.3  103 mmol g1. The degradation rate increased
almost linearly with the extent of adsorption of RhB (Fig. S4B)
and could be adequately described by the LangmuireHinshel-
also provides additional disinfection properties (Liu et al., wood (LeH) equation, r ¼ kLHKCe/(1 D KCe), yielding an LeH rate
2008), however, both activities greatly depend on the amount constant of kLH, ¼ 7.9  104 mM min1 and Langmuir adsorp-
of Ag loaded. As shown in Fig. 5, the photocatalytic degradation tion constant of K ¼ 34.3 mM1. The decrease in RhB adsorption
of RhB and disinfection of E. coli were changed by the loading of with Ag loading would inevitably be expected to lower the rate
Ag nanoparticles with the rate constants of both processes of photocatalytic degradation. Significantly, RhB degradation
higher at low loadings and lower at high loadings. The optimal was much more rapid in an O2 equilibrated suspension than in
loadings of Ag were different with about 0.25 wt% found to be an air saturated solution (Table S1) and was greatly inhibited in
optimal for RhB degradation and about 2.0 wt% optimal for E. an atmosphere of pure N2. Since photocatalytic degradation
coli disinfection. This is perhaps not surprising given that the greatly depends on the amount of ROSs, which were produced
photocatalyst itself exhibits bactericidal properties with the by electron transfer between excited TiO2 and the adsorbed O2
extent of inhibition (in the dark) increasing with increasing Ag (Hoffmann et al., 1995; Sun and Xu, 2009), increasing the
loading while the degradation of RhB is strictly light depen- concentration of O2 presumably greatly increases the rate of
dent. Further discussion of the photocatalytically-mediated production of ROSs, thus enhancing the degradation of RhB. O2
degradation of RhB is provided below. Fig. S3 shows the pho- adsorptionedesorption isotherms obtained through solid-gas
toluminescence spectra of TiO2 with different Ag loadings. TiO2 phase studies (Fig. S5) indicate that as Ag loading increased
has a strong luminescence spectrum with the luminescence from 0 to 3.0%, the BET surface area gradually increased from
intensity decreasing progressively on increasing the Ag loading 134 to 156 m2/g (Fig. 6) and the pore volume increased by about
from 0 to 2.0% (Fig. 5). This effect may be attributed to Schottky 11%. Apparently, the increased affinity of the catalyst to O2
barrier formation at the TiO2emetal interface, indicating could be attributed to the deposition of Ag nanoparticles, which
a strong inhibition of eehþ recombination as reported for increased the adsorption of oxygen on the photocatalyst
other noble metals (Bannat et al., 2009; Sadeghi et al., 1996; surface (Fig. 7) (Xin et al., 2005). As such, two factors may be
Subramanian et al., 2001; Tada et al., 2009; Xin et al., 2005), considered to account for the promotion of RhB degradation.
eehþ recombination was decreased at the low loadings and First, electrons are likely to be transferred from the TiO2 surface
enhanced at the high loadings with an optimal loading of Ag at to the Ag particles due to formation of the Schottky barrier at the
about 0.10%. However, the optimal Ag loading for the interface of TiO2 and Ag (Fig. S3) with the adsorption of O2 on the

Fig. 7 e Schematic illustration of the effect of Ag on the adsorption of RhB and O2.
w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3 2101

4. Conclusions

Mesoporous anatase TiO2 materials modified with Ag nano-


particles displayed significant capacity for the degradation of
RhB and disinfection of E. coli under UV light irradiation. Both
dye degradation and E. coli inactivation increased at low Ag
loadings and decreased at high loadings with optimal Ag load-
ings of about 0.25 and 2.0 wt% for RhB degradation and E. coli
disinfection, respectively. Besides formation of the Schottky
barrier at the TiO2emetal interface, the deposited Ag nano-
particles slightly decreased the extent of adsorption of RhB, but
increased the affinity of the surface to oxygen, which seemed to
play a much more important role in enhancing the photocata-
lytic reactions. After four cycles of reuse, the composite catalyst
still showed significant capacity for dye degradation.
Fig. 8 e Reuse experiment for RhB degradation over
m-TiO2eAg 0.25%. The hollow symbols represent the
concentration of RhB before adsorption. Acknowledgment

This work was supported by Environment and Water Industry


Development Council (EWI) of Singapore under project
Ag surface facilitating the electron transfer process. Second, number MEWR 651/06/161.
since the reduced O2 will transform to various ROSs, the
enhancement of O2 adsorption would generate more ROSs in Appendix. Supporting information.
solution, leading to enhanced RhB degradation. Thus, the
impact of the Ag particles on the photocatalytic activity can be TEM images, Ag 3d XPS spectra, and PL spectra of m-TiO2eAg
summarized as follows: (1) the Ag particles created a Schottky nanocomposite; Sorption isotherms of RhB and O2, and the
barrier that depressed the eehþ recombination and promoted effect of catalyst loading on the photoactivity.
the utilization of excitons, (2) the Ag particles decreased the
adsorption of RhB on the catalyst surface, resulting in a decline
of the photoactivity, (3) the Ag particles increased the adsorp- Appendix. Supplementary data
tion affinity of the catalyst towards dissolved oxygen, leading to
an enhanced production of ROSs and the RhB degradation rate Supplementary data associated with this article can be found
therefore. However, compared to the decrease in RhB adsorp- in the online version at doi: doi:10.1016/j.watres.2010.12.019.
tion with increase in Ag loadings, the incremental increase in
extent of O2 adsorption appeared to play a much more impor-
tant role in the photocatalytic reactions. Thus, the optimal Ag references
loading further increased to 0.25 wt% as compared to sample m-
TiO2eAg 0.10% with the lowest eehþ recombination.
Arabatzis, I.M., Stergiopoulos, T., Bernard, M.C., Labou, D.,
3.4. Stability of photocatalyst Neophytides, S.G., Falaras, P., 2003. Silver-modified titanium
dioxide thin films for efficient photodegradation of methyl
orange. Appl. Catal. B. 42, 187e201.
The recyclability of the catalyst m-TiOeAg 0.25% is shown in Bannat, I., Wessels, K., Oekermann, T., Rathousky, J.,
Fig. 8. Since the photocatalytic inactivation of E. coli follows Bahnemann, D., Wark, M., 2009. Improving the photocatalytic
a similar trend as RhB degradation (Fig. 4), stability of the performance of mesoporous titania films by modification with
catalyst was evaluated with RhB degradation. It can be seen gold nanostructures. Chem. Mater. 21, 1645e1653.
that the photocatalyst was recyclable for continuous adsorp- Beyers, E., Cool, P., Vansant, E.F., 2005. Anatase formation during
tion and removal of RhB under UV irradiation, although the the synthesis of mesoporous titania and its photocatalytic
effect. J. Phys. Chem. B. 109, 10081e10086.
photoactivity and sorption capacity gradually decreased from
Chen, D.H., Huang, F.Z., Cheng, Y.B., Caruso, R.A., 2009a. Mesoporous
one run to another. The results can be attributed to two anatase TiO2 beads with high surface areas and controllable pore
reasons. First, any intermediates formed would be expected to sizes: a superior candidate for high-performance dye-sensitized
compete for the adsorption sites and reactive oxygen species solar cells. Adv. Mater. 21, 2206e2207.
thereby diminishing oxidation effectiveness. Second, the Chen, F.N., Yang, X.D., Wu, Q., 2009b. Photocatalytic oxidation of
catalyst concentration was decreased from one run to another Escherischia coli, aspergillus niger, and formaldehyde under
different ultraviolet irradiation conditions. Environ. Sci.
because of the removal of 5 mL aliquots for analysis purposes.
Technol. 43, 4606e4611.
It is supported by the results shown in Fig. S6. It is seen that
Chen, F.N., Yang, X.D., Xu, F.F., Wu, Q., Zhang, Y.P., 2009c.
the rate constants of both RhB degradation and E. coli inacti- Correlation of photocatalytic bactericidal effect and organic
vation increased with the catalyst loading under the other- matter degradation of TiO2 part I: observation of phenomena.
wise identical conditions to that of Fig. 4. Environ. Sci. Technol. 43, 1180e1184.
2102 w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3

Chen, X., Mao, S.S., 2007. Titanium dioxide nanomaterials: Raveendran, P., Eswaramoorthy, M., Bindu, U., Chatterjee, M.,
synthesis, properties, modifications, and applications. Chem. Hakuta, Y., Kawanami, H., Mizukami, F., 2008. Template-free
Rev. 107, 2891e2959. formation of meso-structured anatase TiO2 with spherical
Epling, W.S., Peden, C.H.F., Henderson, M.A., Diebold, U., 1998. morphology. J. Phys. Chem. C 112, 20007e20011.
Evidence for oxygen adatoms on TiO2(110) resulting from O-2 Sadeghi, M., Liu, W., Zhang, T.G., Stavropoulos, P., Levy, B., 1996.
dissociation at vacancy sites. Surf. Sci. 412e13, 333e343. Role of photoinduced charge carrier separation distance in
Guin, D., Manorama, S.V., Latha, J.N.L., Singh, S., 2007. heterogeneous photocatalysis: oxidative degradation of
Photoreduction of silver on bare and colloidal TiO2 CH3OH vapor in contact with Pt/TiO2 and cofumed
nanoparticles/nanotubes: synthesis, characterization, and TiO2eFe2O3. J. Phys. Chem. 100, 19466e19474.
tested for antibacterial outcome. J. Phys. Chem. C 111, Schattka, J.H., Shchukin, D.G., Jia, J.G., Antonietti, M., Caruso, R.A.,
13393e13397. 2002. Photocatalytic activities of porous titania and titania/
Han, F., Kambala, V.S.R., Srinivasan, M., Rajarathnam, D., zirconia structures formed by using a polymer gel templating.
Naidu, R., 2009. Tailored titanium dioxide photocatalysts for Chem. Mater. 14, 5103e5108.
the degradation of organic dyes in wastewater treatment: Sharma, V.K., Yngard, R.A., Lin, Y., 2009. Silver nanoparticles:
a review. Appl. Catal. A 359, 25e40. green synthesis and their antimicrobial activities. Adv. Colloid
He, J.H., Ichinose, I., Kunitake, T., Nakao, A., 2002. In situ Interfac. 145, 83e96.
synthesis of noble metal nanoparticles in ultrathin TiO2-gel Sing, K.S.W., Everett, D.H., Haul, R.A.W., Moscou, L.,
films by a combination of ion-exchange and reduction Pierotti, R.A., Rouquerol, J., Siemieniewska, T., 1985.
processes. Langmuir 18, 10005e10010. Reporting physisorption data for gas solid systems with
Hidaka, H., Yamada, S., Suenaga, S., Zhao, J., Serpone, N., special reference to the determination of surface-area and
Pelizzetti, E., 1990. Photodegradation of surfactants. 6. porosity (recommendations 1984). Pure Appl. Chem. 57,
Complete photocatalytic degradation of anionic, cationic and 603e619.
nonionic surfactants in aqueous semiconductor dispersions. J. Subramanian, V., Wolf, E., Kamat, P.V., 2001. Semiconductor-
Mol. Catal. 59, 279e290. metal composite nanostructures. To what extent do metal
Hoffmann, M.R., Martin, S.T., Choi, W.Y., Bahnemann, D.W., 1995. nanoparticles improve the photocatalytic activity of TiO2
Environmental applications of semiconductor photocatalysis. films? J. Phys. Chem. B. 105, 11439e11446.
Chem. Rev. 95, 69e96. Sun, Q., Xu, Y.M., 2009. Sensitization of TiO2 with aluminum
Iliev, V., Tomova, D., Bilyarska, L., Eliyas, A., Petrov, L., 2006. phthalocyanine: factors influencing the efficiency for
Photocatalytic properties of TiO2 modified with platinum and chlorophenol degradation in water under visible light. J. Phys.
silver nanoparticles in the degradation of oxalic acid in Chem. C 113, 12387e12394.
aqueous solution. Appl. Catal. B. 63, 266e271. Sunada, K., Watanabe, T., Hashimoto, K., 2003a. Bactericidal
Ismail, A.A., Bahnemann, D.W., Bannat, I., Wark, M., 2009. Gold activity of copper-deposited TiO2 thin film under weak UV
nanoparticles on mesoporous interparticle networks of light illumination. Environ. Sci. Technol. 37, 4785e4789.
titanium dioxide nanocrystals for enhanced photonic Sunada, K., Watanabe, T., Hashimoto, K., 2003b. Studies on
efficiencies. J. Phys. Chem. C 113, 7429e7435. photokilling of bacteria on TiO2 thin film. J. Photochem.
Kang, C., Jing, L., Guo, T., Cui, H., Zhou, J., Fu, H., 2008. Photobiol. A 156, 227e233.
Mesoporous SiO2-modified nanocrystalline TiO2 with high Sung-Suh, H.M., Choi, J.R., Hah, H.J., Koo, S.M., Bae, Y.C., 2004.
anatase thermal stability and large surface area as efficient Comparison of Ag deposition effects on the photocatalytic
photocatalyst. J. Phys. Chem. C 113, 1006e1013. activity of nanoparticulate TiO2 under visible and UV light
Kowalska, E., Remita, H., Colbeau-Justin, C., Hupka, J., Belloni, J., irradiation. J. Photochem. Photobiol. A 163, 37e44.
2008. Modification of titanium dioxide with platinum ions and Tachikawa, T., Fujitsuka, M., Majima, T., 2007. Mechanistic
clusters: application in photocatalysis. J. Phys. Chem. C 112, insight into the TiO2 photocatalytic reactions: design of new
1124e1131. photocatalysts. J. Phys. Chem. C 111, 5259e5275.
Kvitek, L., Panacek, A., Soukupova, J., Kolar, M., Vecerova, R., Tada, H., Kiyonaga, T., Naya, S., 2009. Rational design and
Prucek, R., Holecova, M., Zboril, R., 2008. Effect of surfactants applications of highly efficient reaction systems
and polymers on stability and antibacterial activity of silver photocatalyzed by noble metal nanoparticle-loaded titanium
nanoparticles (NPs). J. Phys. Chem. C 112, 5825e5834. (IV) dioxide. Chem. Soc. Rev. 38, 1849e1858.
Li, H.X., Bian, Z.F., Zhu, J., Huo, Y.N., Li, H., Lu, Y.F., 2007. Tian, B.Z., Yang, H.F., Liu, X.Y., Xie, S.H., Yu, C.Z., Fan, J., Tu, B.,
Mesoporous Au/TiO2 nanocomposites with enhanced Zhao, D.Y., 2002. Fast preparation of highly ordered
photocatalytic activity. J. Am. Chem. Soc. 129, 4538e4539. nonsiliceous mesoporous materials via mixed inorganic
Linsebigler, A.L., Lu, G.Q., Yates, J.T., 1995. Photocatalysis on TiO2 precursors. Chem. Commun., 1824e1825.
surfaces-principles, mechanisms, and selected results. Chem. Tran, H., Scott, J., Chiang, K., Amal, R., 2006. Clarifying the role
Rev. 95, 735e758. of silver deposits on titania for the photocatalytic
Liu, Y., Wang, X.L., Yang, F., Yang, X.R., 2008. Excellent mineralisation of organic compounds. J. Photochem.
antimicrobial properties of mesoporous anatase TiO2 and Ag/ Photobiol. A 183, 41e52.
TiO2 composite films. Micropor. Mesopor. Mater. 114, 431e439. Wang, X., Mitchell, D.R.G., Prince, K., Atanacio, A.J., Caruso, R.A.,
Lv, L., Lu, Y.Q., Ng, W.J., Zhao, X.S., 2009. Bactericidal activity of 2008. Gold nanoparticle incorporation into porous titania
silver nanoparticles supported on microporous titanosilicate networks using an agarose gel templating technique for
ETS-10. Micropor. Mesopor. Mater. 120, 304e309. photocatalytic applications. Chem. Mater. 20, 3917e3926.
Mogyorosi, K., Farkas, A., Dekany, I., Ilisz, I., Dombi, A., 2002. TiO2- Wu, T.X., Liu, G.M., Zhao, J.C., Hidaka, H., Serpone, N., 1998.
based photocatalytic degradation of 2-chlorophenol adsorbed Photoassisted degradation of dye pollutants. V. Self-
on hydrophobic clay. Environ. Sci. Technol. 36, 3618e3624. photosensitized oxidative transformation of rhodamine B
Pan, J.H., Dou, H.Q., Xiong, Z.G., Xu, C., Ma, J.Z., Zhao, X.S., 2010. under visible light irradiation in aqueous TiO2 dispersions.
Porous photocatalysts for advanced water purifications. J. J. Phys. Chem. B. 102, 5845e5851.
Mater. Chem. 20, 4512e4528. Xin, B., Jing, L., Ren, Z., Wang, B., Fu, H., 2005. Effects of
Pinna, N., Niederberger, M., 2008. Surfactant-free nonaqueous simultaneously doped and deposited Ag on the photocatalytic
synthesis of metal oxide nanostructures. Angew. Chem. Int. activity and surface states of TiO2. J. Phys. Chem. B. 109,
Ed. Engl. 47, 5292e5304. 2805e2809.
w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 2 0 9 5 e2 1 0 3 2103

Xu, Y., Lv, K., Xiong, Z., Leng, W., Du, W., Liu, D., Xue, X., 2007. Zhang, F.L., Zheng, Y.H., Cao, Y.N., Chen, C.Q., Zhan, Y.Y., Lin, X.
Rate enhancement and rate inhibition of phenol degradation Y., Zheng, Q., Wei, K.M., Zhu, J.F., 2009. Ordered mesoporous
over irradiated anatase and rutile TiO2 on the addition of NaF: AgeTiO2-KIT-6 heterostructure: synthesis, characterization
new insight into the mechanism. J. Phys. Chem. C 111, and photocatalysis. J. Mater. Chem. 19, 2771e2777.
19024e19032. Zhao, D., Chen, C., Wang, Y., Ma, W., Zhao, J., Rajh, T., Zang, L.,
Xu, Y.M., Langford, C.H., 1995. Enhanced photoactivity of 2008. Enhanced photocatalytic degradation of dye pollutants
a titanium(IV) oxide-supported on ZSM5 and Zeolite-A at low- under visible irradiation on Al(III)-modified TiO2: structure,
coverage. J. Phys. Chem. 99, 11501e11507. interaction, and interfacial electron transfer. Environ. Sci.
Yu, J.C., Zhang, L.Z., Zheng, Z., Zhao, J.C., 2003. Synthesis and Technol. 42, 308e314.
characterization of phosphated mesoporous titanium Zhao, D.Y., Jaroniec, M., Hsiao, B.S., 2010. Editorial for themed
dioxide with high photocatalytic activity. Chem. Mater. 15, issue on “advanced materials in water treatments”. J. Mater.
2280e2286. Chem. 20, 4476e4477.
Zhang, D.Q., Li, G.S., Yu, J.C., 2010. Inorganic materials for Zhou, J.K., Zhang, Y.X., Zhao, X.S., Ray, A.K., 2006.
photocatalytic water disinfection. J. Mater. Chem. 20, Photodegradation of benzoic acid over metal-doped TiO2. Ind.
4529e4536. Eng. Chem. Res. 45, 3503e3511.

You might also like