Top Banner
Nanoporous Au–TiMCM-41An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange Panneer Selvam Sathish Kumar, Michael Ruby Raj, Sambandam Anandan n Nanomaterials & Solar Energy Conversion Lab, Department of Chemistry, National Institute of Technology, Trichy 620 015, India article info Article history: Received 1 October 2009 Accepted 22 May 2010 Available online 9 June 2010 Keywords: Au–TiMCM-41 Nanocomposite MeOr Electron acceptor Photocatalytic degradation Visible light abstract Modification of MCM-41 nanopore materials was found to improve the catalyst activity because it would provide an effective environment for increasing the number of active surface sites for molecular interaction. Hence in this article, with an attempt to extend light absorption of TiMCM-41 based photocatalyst toward the visible light range and to eliminate the rapid recombination of excited electrons/holes during photoreaction, a new type of photocatalyst Au/TiMCM-41 was prepared. The crystal phase composition and surface morphology of the Au/TiMCM-41 photocatalyst were comprehensively examined by X-ray differential detection and transmission electron microscopy analysis. The analytical results of UV–visible diffused reflectance spectra indicated that the absorbance in the visible region above 500 nm for Au/TiMCM-41 photocatalyst made it possible to be excited by visible light. The experiment demonstrated that the photocatalytic degradation of methyl orange using Au/TiMCM-41 photocatalyst was significantly higher than that using TiMCM-41 photocatalyst. Possibilities to enhance catalyst performance further by using electron acceptors are also discussed. & 2010 Elsevier B.V. All rights reserved. 1. Introduction Nanoporous materials play an important role in many aspects of science and technology, of which molecular recognition is becoming increasingly important for environmental protection [1]. Extremely specific molecular selectivity is required for the removal of highly diluted pollutants in water; otherwise many coexisting compounds with higher concentrations will saturate the adsorption capacity [2]. So, in recent years the intense worldwide activity in applications of newly developed simple geometry templated porous media (MCM-41) in which the pore diameter can be accurately tuned in the range from 2 to 25 nm by Mobil researchers [3,4] has led to renewed interest in this subject. However, pure MCM-41 showed very limited catalytic activity due to the lack of lattice defect, the redox properties, the basicity and acidity [5]. The introduction of guests (transition metals) into nanoporous silicates may increase the active sites and thus improve the catalytic activity [6–9]. This is because in MCM-41, pores are not interconnected and the surface contains hydroxyl groups, which may experience specific interactions with mole- cules susceptible of having hydrogen bonds. So, specific surface interaction effects on the properties of the confined compounds may be interest to everyone. To these ends, TiO 2 is the most widely used photocatalyst in the area of photocatalysis [10]. However, TiO 2 has its own drawbacks like fast recombination of photoinduced charge carriers (electrons and holes) and it can only be excited by ultraviolet irradiation ( r365 nm; Eg ¼ 3.2 eV). In order to avoid these drawbacks of using bare TiO 2 as a photocatalytic material, it is necessary to substitute titania into the structure of MCM-41 nanochannels, a member of new family of nanoporous materials may lead to new catalysts. These hybrid nanocrystalline nanopor- ous materials are expected to show very high active surface area and enhanced photocatalytic efficiencies [11]. Earlier studies showed that MCM-41 materials can achieve stable photoinduced charge separation [12,13]. Further, incorporation of reducible transition metal ions into nanoporous materials impedes back electron transfer by acting as more stable electron acceptors [12–15]. Gold was recognized to be poorly active as a catalyst; however, when Au nanoparticles are highly dispersed on semiconductor metal oxides or hydroxides, they exhibit good catalytic activity. Such unusually high activity of gold strongly interacting with the support has later been confirmed by many research groups [16–20]. Further, it is expected that the encapsulation of gold nanoparticles into the MCM-41/TiMCM-41 surface can tailor the photoresponsiveness into the visible region ( r400 nm). Hence in the present work, we have studied the influence of gold deposition in the nanoporous TiMCM-41 (prepared by two modes of incorporation of Ti into MCM-41 nanochannel) for the visible light assisted photocatalytic degradation of azo dye Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/solmat Solar Energy Materials & Solar Cells 0927-0248/$ - see front matter & 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.solmat.2010.05.046 n Corresponding author. Tel.: + 91 44 2503639; fax: + 91 431 2500133. E-mail addresses: [email protected], [email protected] (S. Anandan). Solar Energy Materials & Solar Cells 94 (2010) 1783–1789
7

Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

Apr 28, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

Solar Energy Materials & Solar Cells 94 (2010) 1783–1789

Contents lists available at ScienceDirect

Solar Energy Materials & Solar Cells

0927-02

doi:10.1

n Corr

E-m

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

Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst towardvisible photooxidation of methyl orange

Panneer Selvam Sathish Kumar, Michael Ruby Raj, Sambandam Anandan n

Nanomaterials & Solar Energy Conversion Lab, Department of Chemistry, National Institute of Technology, Trichy 620 015, India

a r t i c l e i n f o

Article history:

Received 1 October 2009

Accepted 22 May 2010Available online 9 June 2010

Keywords:

Au–TiMCM-41

Nanocomposite

MeOr

Electron acceptor

Photocatalytic degradation

Visible light

48/$ - see front matter & 2010 Elsevier B.V. A

016/j.solmat.2010.05.046

esponding author. Tel.: +91 44 2503639; fax

ail addresses: [email protected], sanand@

a b s t r a c t

Modification of MCM-41 nanopore materials was found to improve the catalyst activity because it

would provide an effective environment for increasing the number of active surface sites for molecular

interaction. Hence in this article, with an attempt to extend light absorption of TiMCM-41 based

photocatalyst toward the visible light range and to eliminate the rapid recombination of excited

electrons/holes during photoreaction, a new type of photocatalyst Au/TiMCM-41 was prepared. The

crystal phase composition and surface morphology of the Au/TiMCM-41 photocatalyst were

comprehensively examined by X-ray differential detection and transmission electron microscopy

analysis. The analytical results of UV–visible diffused reflectance spectra indicated that the absorbance

in the visible region above 500 nm for Au/TiMCM-41 photocatalyst made it possible to be excited by

visible light. The experiment demonstrated that the photocatalytic degradation of methyl orange using

Au/TiMCM-41 photocatalyst was significantly higher than that using TiMCM-41 photocatalyst.

Possibilities to enhance catalyst performance further by using electron acceptors are also discussed.

& 2010 Elsevier B.V. All rights reserved.

1. Introduction

Nanoporous materials play an important role in many aspectsof science and technology, of which molecular recognition isbecoming increasingly important for environmental protection [1].Extremely specific molecular selectivity is required for theremoval of highly diluted pollutants in water; otherwise manycoexisting compounds with higher concentrations will saturatethe adsorption capacity [2]. So, in recent years the intenseworldwide activity in applications of newly developed simplegeometry templated porous media (MCM-41) in which the porediameter can be accurately tuned in the range from 2 to 25 nm byMobil researchers [3,4] has led to renewed interest in this subject.However, pure MCM-41 showed very limited catalytic activity dueto the lack of lattice defect, the redox properties, the basicity andacidity [5]. The introduction of guests (transition metals) intonanoporous silicates may increase the active sites and thusimprove the catalytic activity [6–9]. This is because in MCM-41,pores are not interconnected and the surface contains hydroxylgroups, which may experience specific interactions with mole-cules susceptible of having hydrogen bonds. So, specific surfaceinteraction effects on the properties of the confined compoundsmay be interest to everyone.

ll rights reserved.

: +91 431 2500133.

nitt.edu (S. Anandan).

To these ends, TiO2 is the most widely used photocatalyst inthe area of photocatalysis [10]. However, TiO2 has its owndrawbacks like fast recombination of photoinduced chargecarriers (electrons and holes) and it can only be excited byultraviolet irradiation (r365 nm; Eg¼3.2 eV). In order to avoidthese drawbacks of using bare TiO2 as a photocatalytic material, itis necessary to substitute titania into the structure of MCM-41nanochannels, a member of new family of nanoporous materialsmay lead to new catalysts. These hybrid nanocrystalline nanopor-ous materials are expected to show very high active surfacearea and enhanced photocatalytic efficiencies [11]. Earlier studiesshowed that MCM-41 materials can achieve stable photoinducedcharge separation [12,13]. Further, incorporation of reducibletransition metal ions into nanoporous materials impedesback electron transfer by acting as more stable electron acceptors[12–15].

Gold was recognized to be poorly active as a catalyst; however,when Au nanoparticles are highly dispersed on semiconductormetal oxides or hydroxides, they exhibit good catalytic activity.Such unusually high activity of gold strongly interacting withthe support has later been confirmed by many research groups[16–20]. Further, it is expected that the encapsulation of goldnanoparticles into the MCM-41/TiMCM-41 surface can tailor thephotoresponsiveness into the visible region (r400 nm).

Hence in the present work, we have studied the influence ofgold deposition in the nanoporous TiMCM-41 (prepared by twomodes of incorporation of Ti into MCM-41 nanochannel) for thevisible light assisted photocatalytic degradation of azo dye

Page 2: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–17891784

(methyl orange). The gold nanoparticles were introduced into theTiMCM-41 nanochannels through the deposition-precipitationmethod (�2 atomic weight %). In addition to this externalelectron acceptors such as peroxomonosulphate (PMS), perox-odisulphate (PDS) and hydrogen peroxide (H2O2) were added andtheir superiority in degrading the target pollutant was observed inorder to get effective outcome in the field of environmentalresearch.

2. Experimental

2.1. Materials

Chloroauric acid trihydrate (HAuCl4 �3H2O), hexadecyltri-methylammonium bromide (CTAB), tetraethyl orthosilicate(TEOS), titanium isopropoxide and ammonium hydroxide(NH4OH) were purchased from Sigma-Aldrich and used as astarting material to prepare Au–TiMCM-41 nanocomposites.Methyl orange (MeOr; dye content 85%) was purchased fromSigma-Aldrich and used without further purification. Potassiumperoxomonosulphate, a triple salt with the composition of2KHSO5 �KHSO4 �K2SO4 (commercially known as ‘‘Oxone’’) fromJanssen Chimica (Belgium), was used as received. Unless other-wise specified, all the reagents used were of analytical grade andthe solutions were prepared using millipore water.

2.2. Preparation of the photocatalyst

2.2.1. Preparation of TiMCM-41 by hydrothermal crystallization

method

TiMCM-41 hybrid material was prepared by the hydrothermalcrystallization method [21] as follows: about 2.2 g of CTAB wasdissolved in 52 ml of water at 40 1C. 26 ml of ammoniumhydroxide (28%) was then added slowly under stirring. Additionof 10 ml of TEOS and 0.03 g of titanium isopropoxide to thesolution was followed by continued stirring for 3 h at roomtemperature. The resulting gel was then transferred into a Teflon-lined autoclave and set aside for 48 h at 110 1C. Filtration andwashing with distilled water until the mother liquor reaches theneutral pH gave TiMCM-41. Template removal was typicallyconducted by heating at 100 1C for 12 h followed by calcination at550 1C for 5 h. Similarly, MCM-41 is prepared without titaniumisopropoxide for comparison.

2.2.2. Preparation of TiMCM-41 by grafting technique

As-synthesized MCM-41 materials were dehydrated invacuum (10�3 Torr) for 48 h. Then titanium grafting wasperformed as follows: 0.5 g of the vacuum-dehydrated MCM-41sample was added into a solution of 0.045 g titanium isoprop-oxide in 10 g of anhydrous hexane under vigorous stirring in N2

atmosphere. The mixture stood for 1 h and was then filtered andwashed with anhydrous hexane. The grafted material wascalcined at 550 1C for 5 h in oxygen atmosphere to convertunreacted alkoxide ligands into Ti–OH groups and to removeresidual isopropyl alcohol and hexane.

2.2.3. Deposition of gold nanoparticles at the TiMCM-41

nanochannels

Encapsulation of gold nanoparticles at the TiMCM-41 nano-channels were performed by deposition-precipitation methodwith NaOH [22,23] as follows: 100 ml of an aqueous solution ofHAuCl4 �3H2O (1.05�10�3 M) was heated to 80 1C. The pH of thesolution was adjusted to 7 by dropwise addition of NaOH (1 M),then 1 g of TiMCM-41 was dispersed in the solution and the pH of

the suspension was readjusted to 7 by the dropwise addition ofNaOH (1 M). The suspension was then thermostated at 80 1C withvigorous stirring for 2 h and the solids were gathered bycentrifugation (12,000 rpm for 10 min) and washed with 100 mlof distilled water under stirring for 10 min at 50 1C. The aboveprocedure was repeated several times and the samples were driedunder vacuum at 100 1C for 2 h. The final yield of gold loading was�2 atomic weight % on TiMCM-41.

2.3. Characterization techniques

Surface morphology, particle size and the various contours ofthe photocatalyst powders were analyzed by XRD (measuredusing Rigaku diffractometer, Cu-Ka radiation, Japan) and trans-mission electron microscopy (recorded using TECNAI G2 model),respectively. Diffuse reflectance UV–vis spectra of the sampleswere recorded using a Shimadzu UV–vis 2550 spectrophotometer.The surface area, pore volume and pore diameter of the samplewere measured with the assistance of Flowsorb II 2300 ofMicrometrics, Inc. Infrared spectra were recorded using aShimadzu FT-IR 8400 S spectrophotometer by employing KBrpellet technique.

2.4. Evaluation of photocatalytic activity

The photocatalytic experiments were conducted under ambi-ent atmospheric conditions and at natural pH (�6.0) using 150 Wtungsten halogen lamp (lZ400 nm; intensity of incident radia-tion is 80600710 Lux measured using Extec, USA) as the lightsource. In order to ensure adsorption/desorption equilibrium, thesolution was stirred for about 45 min in dark, prior to irradiation.The apparent kinetics of disappearance of the substrate, MeOr,was determined by following the concentration of the substrate(lmax¼466 nm) using a UV–vis spectrophotometer (PG Instru-ments, UK) after a certain period of irradiation of the photo-catalyst suspension. Prior to the analysis, the solid catalyst wasremoved from samples by filtration through a 0.45 mm poly-vinylidene fluoride (PVDF) filter.

3. Results and discussion

3.1. Characterization of the photocatalyst

The surface morphology of the as-prepared Au/TiMCM-41nanocatalyst was identified from transmission electron microscopy(TEM). The ordered hexagonal pore arrangements of pure MCM-41and TiMCM-41 prepared via hydrothermal method and graftingmethod are clearly visible from Fig. 1(A)–(C). Fig. 1(D)–(F) clearlydepicts that the Au nanoparticles were found to be uniformlydispersed on the support surfaces without affecting the MCM-41morphology [24]. The corresponding selected area electrondiffraction (SAED) pattern is shown in the inset of Fig. 1(E) and (F)confirming the formation of the single crystalline Au nanoparticlesas well. The size of the as-prepared gold nanoparticles was 5–15 nm.

Upon formation of MCM-41 by the condensation reactionsbetween the surfactant and organoalkoxysilane the followingtrend is generally observed with respect to nanochannels: that is,pore sizes of the channels are significantly reduced or at the sametime it is expected that the wall thickness may be increased[25,26]. Hence, it is important to measure the pore diameter, porevolume and specific surface area of all the prepared samplesusing nitrogen adsorption/desorption isotherm and multi-pointBET analysis [27,28]. The observed results for the preparedphotocatalysts were tabulated in Table 1. The surface area of

Page 3: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

Fig. 1. High resolution TEM micrograph of unmodified MCM-41 (A), titania modified MCM-41 via hydrothermal method (B), titania modified MCM-41 via grafting method

(C), gold supported titania modified MCM-41 via hydrothermal method (D, E) and gold supported titania modified MCM-41 via grafting method (F). Inset of E and F shows

SAED pattern of gold supported titania modified MCM-41 via hydrothermal and grafting methods.

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–1789 1785

as-prepared MCM-41 is 1029 m2/g and the pore diameter is3.0 nm. However for encapsulation of titania into MCM-41 adecrease in surface area and pore volume was observed either

prepared via hydrothermal method (911 m2/g; 2.82 nm) orgrafting method (938 m2/g; 2.90 nm). Further deposition of goldnanoparticles, surface area and pore volume are reduced for both

Page 4: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

Table 1Physicochemical characteristics of modified photocatalysts.

Catalyst Surface area [SBET] (m2 g�1) Pore volume (cm3 g�1) Pore diameter (nm)

MCM-41 1029 0.72 3.00

Ti-MCM-41 (hydrothermal method) 911 0.64 2.82

Ti-MCM-41 (grafting method) 938 0.66 2.90

Au/Ti-MCM-41 (hydrothermal method) 825 0.60 2.62

Au/Ti–MCM-41 (grafting method) 771 0.62 2.78

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

6000

7000

8000

9000

10000

11000

E

D

C

Inte

nsity

(arb

. uni

ts)

2 (θ) degree

TiMCM-41 (Grafting method) Au/TiMCM-41 (Hydrothermal method) Au/TiMCM-41 (Grafting method)

0 2 4 6 8 10 12 14

0

2000

4000

6000

8000

10000

B

A

Inte

nsity

(arb

. uni

ts)

2 (θ) degree

MCM-41 Ti-MCM-41 (Hydrothermal method)

Fig. 2. XRD pattern of unmodified MCM-41 (A), titania modified MCM-41 via hydrothermal method (B), titania modified MCM-41 via grafting method (C), gold supported

titania modified MCM-41 via hydrothermal method (D) and gold supported titania modified MCM-41 via grafting method (E).

200 300 400 500 600 700 800

0.1

0.2

0.3

0.4

0.5

D

C

B

A

Abs

orba

nce

Wavelength (nm)

Fig. 3. Diffuse reflectance UV–vis spectra of titania modified MCM-41 via

hydrothermal method (A), titania modified MCM-41 via grafting method (B), gold

supported titania modified MCM-41 via hydrothermal method (C) and gold

supported titania modified MCM-41 via grafting method (D).

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–17891786

the samples (see Table 1). This indicates that pore volumedecreases upon addition of transition metals or in other wordsthickness of pore wall increases.

Fig. 2 shows the X-ray diffraction pattern of Au–TiMCM-41,TiMCM-41 and MCM-41 samples at low angle measurements. Onecan easily find XRD peaks at low angle 1.8–101, which indicate thewell ordered nanostructure of MCM-41 is retained even after theinsertion of titania and gold nanoparticles into the MCM-41framework. Moreover additional peak at a higher angle of 251indicates the characteristic of titania and it is present in theanatase form (80%) at the nanochannels [29] (figure not given).However characteristic of gold peak at higher angle (38.41 and44.61) may overlap with MCM-41 peaks and it is not visualizedclearly in XRD patterns (figure not given).

Fig. 3 displays the diffuse reflectance UV–vis spectra ofAu–TiMCM-41, and TiMCM-41. For Au–TiMCM-41, the absorptionband centered at about �550 nm may be due to the attribution ofsurface plasmon resonance of gold nanoparticles. TiMCM-41samples show a broad peak in the UV region, which may be dueto the presence of titania in tetrahedral coordination [30,31] inMCM-41 nanochannels.

Fig. 4 shows FT-IR spectra of the as-prepared Au–TiMCM-41,TiMCM-41 and MCM-41 samples. All the samples show bands at

Page 5: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

1400 1200 1000 800 600 400

E

D

C

B

A

% T

rans

mitt

ance

Wave number ( cm-1 )

Fig. 4. FT-IR spectra of unmodified MCM-41 (A), titania modified MCM-41 via

hydrothermal method (B), titania modified MCM-41 via grafting method (C), gold

supported titania modified MCM-41 via hydrothermal method (D) and gold

supported titania modified MCM-41 via grafting method (E).

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50 60 70Time (min)

-ln (

C/C

0)

EAB

C

D

Fig. 5. Plot of photodegradation of MeOr in the presence of titania modified

MCM-41 via hydrothermal method (A), titania modified MCM-41 via grafting

method (B), gold supported titania modified MCM-41 via hydrothermal method

(C), gold supported titania modified MCM-41 via grafting method (D) and in the

presence of colloidal TiO2 (E). CMeOr¼5�10�5 M; CCat¼1 g L�1.

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–1789 1787

1080 and 810 cm�1 correspond to asymmetric and symmetricSi–O stretching vibrations. The bands at 963 cm�1 were due tothe stretching and bending vibration of surface Si–O� groups.Further, this is also indicative of the formation of Si–O–Ti linkage.As the heteroatoms were introduced to MCM-41, a slight red shiftwas observed for almost all the bands. It may be because thereis a strong interaction between the atoms of gold, titaniumand silicon, i.e., the heteroatoms were incorporated into theframework of MCM-41 [32].

4. Photocatalytic degradation of methyl orange

The catalytic properties of Au–TiO2 and related systemshave attracted chemical industries [33–35] as well as academia[36–39]. But no work has been reported on removal of dyes usingAu–TiMCM-41 nanoporous materials. Hence we studied herephotocatalytic degradation of methyl orange using Au–TiMCM-41nanoporous material and visible light. After checking that there isno appreciable degradation of methyl orange with visible lightirradiation alone, the photocatalytic degradation of methyl orangewas studied in the presence of Au–TiMCM-41 nanoporousmaterials. But for unmodified MCM-41, no photodegradation ratewas observed.

First, the effect of Au–TiMCM-41 (1.0 g L�1) on the photo-catalytic degradation of MeOr (5�10�5 M) was studied inaqueous solution (natural pH; �6.0) and the results werecompared with respect to the same amount of TiMCM-41. Anincrease in � ln(C/Co) corresponds to a decrease in the concentra-tion of the dye. It can be seen that the degradation of MeOroccurred under irradiation in the presence of Au–TiMCM-41. Thelinear relationship between � ln(C/Co) vs irradiation time withrespect to Au–TiMCM-41 indicates that the photocatalyticdegradation follows a pseudo first-order kinetics (Fig. 5). Theobserved rate constant for Au–TiMCM-41 (Table 2) wassignificantly higher compared to that of TiMCM-41 as well aswith that of pure colloidal TiO2. Modification of MCM-41nanopore materials, i.e., Au on titanosilicate supports, was foundto improve the catalyst activity because it would providean effective environment for increasing the number of activesurface sites for dye–semiconductor interaction, which in turn

increases the photodegradation processes. Further, the reason forenhanced efficiency may be due to the following cooperativeeffects: (i) enhanced adsorption of MeOr on the Au–TiMCM-41surface, leading to more injection of the photoexcited electronfrom MeOrn to the conduction band of semiconductor in self-photosensitization pathway under visible light irradiation[40–42]; (ii) the surface plasmon resonance of Au nanoparticleson TiMCM-41 is excited by visible light, enhancing the surfaceelectron excitation and electron–hole separation [43–45]; (iii) Aunanoparticles acting as electron traps to impede electron/holerecombination [46,47]; and (iv) Fermi level equilibration betweenAu nanoparticles and semiconductor may decrease the band gapof the semiconductor and in turn diminish the rapidrecombination of electron–hole pairs [48–50].

Further, it is also known that the photocatalytic activities arelimited by the recombination of photogenerated hole–electronpairs [46]. Hence, in order to improve the efficiencies ofphotodegradation of MeOr, experiments were carried out in thepresence of external electron acceptors (PMS, PDS, H2O2). Further,the use of these electron acceptors alone in the absence of thephotocatalyst did not degrade the methyl orange. In order tocompare the photocatalytic activities, experiments were per-formed using a fixed amount of Au–TiMCM-41 (1.0 g L�1),MeOr (5�10�5 M) and electron acceptors (8�10�5 M). ForAu–TiMCM-41 prepared by the grafting method, a completedecolorization (100%) of the dye was observed in the presence ofPMS (k¼21.27�10�4 s�1), 3.9% of decolorization was achievedwith PDS (k¼0.82�10�4 s�1) and 1.07% of decolorization wasachieved with H2O2 (k¼0.229�10�4 s�1) while only 0.6% ofdecolorization was effected in the absence of any oxidants in60 min, under the same experimental conditions as depictedin Fig. 6. Compared to the rate constant observed in the absenceof the electron acceptors, about 200-fold increase in therate constant is observed for Au–TiMCM-41 nanocompositesin the presence of PMS. The reason for the increase in thephotodegradation rate may be due to the immediate trapping ofthe photogenerated electrons by electron acceptors, which in turndecreases the recombination of electron–hole pairs [51]. Similartrend is also observed in the case of Au–TiMCM-41 prepared bythe hydrothermal method. This is because the distance of theperoxo bond (–O–O–) in electron acceptors is comparable amongthe symmetrical peroxides (H2O2, PDS) and non-symmetricalperoxide (PMS), and it is believed that non-symmetrical peroxidemay be more easily activated than the symmetrical peroxides

Page 6: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

Table 2Comparison of photocatalytic degradation rate of MeOr (5�10�5 M) using different types of catalyst in the presence and absence of PMS (8�10�5 M). Amount of catalyst

maintained for the experimentis 1 gL�1.

Catalyst [1 g L�1]

Rate Colloidal Ti-MCM-41 Ti-MCM-41 Au–Ti-MCM-41 Au–Ti-MCM-41

constant (s�1) TiO2 (H.M.)a (G.M.)b (H.M.)a (G.M.)b

Without PMS 1.13�10�9 1.34�10�9 1.95�10�9 1.15�10�5 1.43�10�5

With PMS 9.21�10�9 1.71�10�8 1.81�10�8 15.53�10�4 21.27�10�4

a Hydrothermal method.b Grafting method.

With out Electron

acceptors

With PMS

With PDSWith H2O2

0

5

10

15

20

25

kx10

4 s-1

Fig. 6. Comparison of photodegradation of MeOr using gold supported titania

modified MCM-41 prepared via grafting method in the presence and absence of

electron acceptors. Concentrations are maintained as follows: MeOr [5�10�5 M],

PMS¼ PDS¼H2O2 0.08 mM, Au–TiMCM-41 (1.0 g L�1).

A B

C

D

0

5

10

15

20

25

kX

10

s

Fig. 7. Comparison of photodegradation of MeOr in the presence of PMS using

titania modified MCM-41 via hydrothermal method (A), titania modified MCM-41

via grafting method (B), gold supported titania modified MCM-41 via hydro-

thermal method (C) and gold supported titania modified MCM-41 via grafting

method (D). Concentrations are maintained as follows: MeOr (5�10�5 M),

PMS (0.08 mM), TiMCM-41/Au–TiMCM-41 (1.0 g L�1).

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–17891788

[52,53]. Hence PMS is more active than PDS (Eqs. (1)–(3)). But theobserved activity with respect to H2O2 is very low compared toPDS. This may be due to the fact that in H2O2 although theformation of hydroxyl radical is increased, the scavengingreaction of the hydroxyl radicals predominates and may resultin a reduced dye degradation rate (Eqs. (4)–(7)) [15,54]

�O3S–O–O–SO3� +e�-SO4

2� +SO4�� (1)

(2,3)

H2O2+e�-OH�+OH� (4)

H2O2+h+-H+ +HO2� (5)

HO2�+HO2

�-H2O2+O2 (6)

OH�+H2O2-H2O+HO2� (7)

The observed rate constant of photooxidation of methyl orangein the presence of TiMCM-41/Au–TiMCM-41 prepared by graftingmethod is approximately equivalent to TiMCM-41/Au–TiMCM-41prepared by hydrothermal method in the presence and absence ofPMS (Fig. 7; Table 2). This is because modification of the MCM-41surface through either method would provide an effectiveenvironment for increasing the number of surface active sitesfor dye–semiconductor interaction. However, the slight increase

in rate constant with respect to TiMCM-41/Au–TiMCM-41prepared by the grafting method might be due to higher TiO2

concentration used in the preparation procedure in the case ofgrafting method compared to the hydrothermal method.

5. Conclusions

In the present study, we found that an otherwise inactiveMCM-41 or less active Ti-MCM-41 upon modification with Ausupport became a new type of inorganic hybrid effectivephotocatalyst for the visible light photodegradation of methylorange. The results suggest that the photocatalytic efficiency ofTiMCM-41 could be significantly enhanced by Au loading, whichincreased the visible light absorption of the catalyst, decreasedthe recombination of the photogenerated charge carriers, in-creased the electron injection from MeOrn and subsequent chargetrapping by the titania matrix. Thus, a combination of increasedavailability of charge carriers due to enhanced photosensitizationeffect and visible light absorption might have played a major rolein enhancing the efficiency of the photocatalyst. It has also beenshown that the dye degradation efficiency could be furtherenhanced in the presence of external electron acceptors, such asPMS.

Acknowledgements

SA thanks CSIR, New Delhi, for the sanction of major researchfund (CSIR reference no. 01(2197)/07/EMR-II). Mr. Sathish Kumarthanks AICTE, New Delhi, for the NDF fellowship.

Page 7: Nanoporous Au–TiMCM-41—An inorganic hybrid photocatalyst toward visible photooxidation of methyl orange

P.S. Sathish Kumar et al. / Solar Energy Materials & Solar Cells 94 (2010) 1783–1789 1789

References

[1] R. Jackson, Transport In Porous Catalysts, Elsevier, Amsterdam, 1977.[2] K. Inumaru, Y. Inoue, S. Kakii, T. Nakono, S. Yamanaka, Organic–inorganic

cooperative molecular recognition in nanostructure of alkyl-grafted MCM-41,Chem. Lett. 32 (2003) 1110–1111.

[3] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck, Orderedmesoporous molecular sieves synthesized by a liquid crystal templatemechanism, Nature 359 (1992) 710–712.

[4] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T. Kresge, K.D. Schmitt,C.T.W. Chu, D.H. Olson, E.W. Sheppard, S.B. McCullen, J.B. Higgins,J.L. Schlenker, A new family of mesoporous molecular sieves prepared withliquid crystal templates, J. Am. Chem. Soc. 114 (1992) 10834–10843.

[5] Y. Kong, S.Y. Jiang, J. Wang, S.S. Wang, Q.J. Yan, Y.N. Lu, Synthesis andcharacterization of Cu–Ti–MCM41, Microporous Mesoporous Mater. 86(2005) 191–197.

[6] Y.J. Do, J.H. Kim, J.H. Park, S.S. Park, S.S. Hong, C.S. Suh, G.D. Lee, Photocatalyticdecomposition of 4-nitrophenol on Ti-containing MCM-41, Catal. Today 101(2005) 299–305.

[7] Z. Luan, E.M. Maes, P.A.W. Van der Heide, D. Zhao, R.S. Czernuszewicz,L. Kevan, Incorporation of titanium into mesoporous silica molecular sieveSBA-15, Chem. Mater. 11 (1999) 3680–3686.

[8] M. Cheng, Z. Wang, K. Sakurai, F. Kumata, T. Saito, T. Komatsu, T. Yashima,Creation of acid sites on SBA-15 mesoporous silica by alumination, Chem.Lett. 28 (1999) 131–132.

[9] Z. Luan, J.Y. Bae, L. Kevan, Vanadosilicate mesoporous SBA-15 molecularsieves incorporated with N-alkylphenothiazines, Chem. Mater. 12 (2000)3202–3207.

[10] S. Anandan, M. Yoon, Photocatalytic activities of the nano-sized TiO2-supported Y-zeolites, J. Photochem. Photobiol. C Photochem. Rev. 4 (2003)5–18.

[11] A. Bhattacharyya, S. Kawi, M.B. Ray, Photocatalytic degradation of orange IIby TiO2 catalysts supported on adsorbents, Catal. Today 98 (2004) 431–439.

[12] H.M. Sung-Suh, Z. Luan, L. Kevan, Photoionization of porphyrins inmesoporous siliceous MCM-41, AlMCM-41, and TiMCM-41 molecular sieves,J. Phys. Chem. B 101 (1997) 10455–10463.

[13] R.M. Krishna, A.M. Prakash, L. Kevan, Photoionization of N-alkylphenothia-zines in mesoporous SiMCM-41, AlMCM-41, and TiMCM-41 molecular sieves,J. Phys. Chem. B 104 (2000) 1796–1801.

[14] S. Sinlapadech, R.M. Krishna, Z. Luan, L. Kevan, Photoionization ofN-alkylphenothiazines in mesoporous SiMCM-41, AlMCM-41, and TiMCM-41 molecular sieves, J. Phys. Chem. B 105 (2001) 4350–4355.

[15] S. Anandan, Photocatalytic effects of titania supported nanoporous MCM-41on degradation of methyl orange in the presence of electron acceptors, DyesPigm. 76 (2008) 535–541.

[16] M. Haruta, Novel catalysis of gold deposited on metal oxides, Catal. Surv. Jpn.1 (1997) 61–73.

[17] M. Haruta, Adsorption of CO on gold supported on TiO2, Catal. Today 36(1997) 115–123.

[18] D. Andreeva, V. Idakiev, T. Tabakova, A. Andreev, Low-temperature water–gasshift reaction over Au/a-Fe2O3, J. Catal. 158 (1996) 354–355.

[19] D. Andreeva, V. Idakiev, T. Tabakova, A. Andreev, R. Vanoli, Low-temperaturewater-gas shift reaction on Au/a-Fe2O3 catalyst, Appl. Catal. A Gen. 134(1996) 275–283.

[20] D. Andreeva, V. Idakiev, T. Tabakova, R. Giovanoli, Low-temperature water-gas shift reaction on Au/TiO2, Au/a-Fe2O3 and Au/Co3O4, Bulg. Chem.Commun. 30 (1998) 59–67.

[21] W. Lin, H. Han, H. Feri, CO2 Splitting by H2O to CO and O2 under UV light inTiMCM-41 silicate sieve, J. Phys. Chem. B 108 (2004) 18269–18273.

[22] R. Zanella, L. Delannoy, C. Louis, Mechanism of deposition of gold precursorsonto TiO2 during the preparation by cation adsorption and deposition–precipitation with NaOH and urea, Appl. Catal. A Gen. 291 (2005) 62–72.

[23] R. Zanella, S. Giorgio, C.H. Shin, C.R. Henry, C. Louis, Characterization andreactivity in CO oxidation of gold nanoparticles supported on TiO2 preparedby deposition-precipitation with NaOH and urea, J. Catal. 222 (2004) 357–367.

[24] C. Liu, X. Yu, J. Yang, M. He, Preparation of mesoporous Al-MCM-41 withstable tetrahedral aluminum using ionic liquids as a single template, Mater.Lett. 61 (2007) 5261–5264.

[25] M.H. Lim, A. Stein, Comparative studies of grafting and direct syntheses ofinorganic�organic hybrid mesoporous materials, Chem. Mater. 11 (1999)3285–3295.

[26] S.D. Sims, S.L. Burkett, S. Mann, The synthesis and characterization of orderedmesoporous materials incorporating organo-siloxanes, Mater. Res. Soc. Symp.Proc. 431 (1996) 77–82.

[27] V.B. Fenelonov, V.N. Romannikov, A.Y. Derevyankin, Mesopore size andsurface area calculations for hexagonal mesophases (types MCM-41, FSM-16,etc.) using low-angle XRD and adsorption data, Microporous MesoporousMater. 28 (1999) 57–72.

[28] K. Miyasaka, A.V. Neimark, O. Terasaki, Density functional theory of in situsynchrotron powder X-ray diffraction on mesoporous crystals: argonadsorption on MCM-41, J. Phys. Chem. C 113 (2009) 791–794.

[29] R.A. Spurr, H. Myers, Quantitative analysis of anatase-rutile mixtures with anX-ray diffractometer, Anal. Chem. 29 (1957) 760–762.

[30] A. Corma, N.T. Navarro, J. Perez-Pariente, Synthesis of an ultralarge poretitanium silicate isomorphous to MCM-41 and its application as a catalyst forselective oxidation of hydrocarbons, J. Chem. Soc. Chem. Commun. (1994)147–148.

[31] M.D. Alba, Z. Luan, J. Klinowski, Titanosilicate mesoporous molecular sieveMCM-41: synthesis and characterization, J. Phys. Chem. 100 (1996)2178–2182.

[32] B.S. Uphade, N. Yamada, T. Nakamura, M. Haruta, Synthesis and character-ization of Ti-MCM-41 and vapor-phase epoxidation of propylene using H2

and O2 over Au/Ti-MCM-41, Appl. Catal. A Gen. 215 (2001) 137–148.[33] R.G. Bowman, H. W. Clark, J.J. Maj, G.E. Hartwell (1998) PCT/US97/11414, PCT

Pub. No. WO 98/00413.[34] T. Hayashi, M. Wada, M. Haruta, S. Tsubota (1999) Jpn. Pat. Appl. No. H10-

244156, PCT Pub. No. WO97/00869, US Patent 5,932,750.[35] A. Kuperman, R.G. Bowman, H.W. Clark, G.E. Hartwell, B.J. Schoeman,

H.E. Tuinstra, G.R. Meima (2001) US Patent 6,255,499.[36] T.A. Nijhuis, H. Huizinga, M.. Makkee, J.A. Moulijn, Direct epoxidation of

propene using gold dispersed on TS-1 and other titanium-containingsupports, Ind. Eng. Chem. Res. 38 (1999) 884–891.

[37] E.E. Stangland, K.B. Stavens, R.P. Andres, W.N. Delgass, Characterization ofgold–titania catalysts via oxidation of propylene to propylene oxide, J. Catal.191 (2000) 332–347.

[38] G. Mul, A. Zwijnenburg, B. Linden, M. van der Makkee, J.A. Moulijn, Stabilityand selectivity of Au/TiO2 and Au/TiO2/SiO2 catalysts in propene epoxidation:an in situ FT-IR study, J. Catal. 201 (2001) 128–137.

[39] M. Haruta, Gold as a low-temperature oxidation catalyst: factors controllingactivity and selectivity, Stud. Surf. Sci. Catal. 110 (1997) 123–134 andreferences therein.

[40] J.M. Herrmann, J. Disdier, P. Pichat, Oxygen species ionosorbed onpowder photocatalyst oxides from room-temperature photoconductivityas a function of oxygen pressure, J. Chem. Soc. Far. Trans. I 77 (1981)2815–2826.

[41] A. Henglein, Reactions of organic free radicals at colloidal silver in aqueoussolution. Electron pool effect and water decomposition, J. Phys. Chem. 83(1979) 2209–2216.

[42] J.M. Herrmann, J. Disdier, P. Pichat, Photoassisted platinum deposition onTiO2 powder using various platinum complexes, J. Phys. Chem. 90 (1986)6028–6034.

[43] G. Zhao, H. Kouzka, T. Yoko, Sol–gel preparation and photoelectrochemicalproperties of TiO2 films containing Au and Ag metal particles, Thin SolidFilms 277 (1996) 147–154.

[44] J.M. Herrmann, H. Tahiri, Y. Ait-Ichou, G. Lassaletta, A.R. Gonzales-Elipe,A. Fernandez, Characterization and photocatalytic activity in aqueous mediumof TiO2 and Ag–TiO2 coatings on quartz, Appl. Catal. B 13 (1997) 219–226.

[45] G. Lassaletta, A.R. Gonzales-Elipe, A. Justo, A. Fernandez, F.J. Ager,M.A. Respaldiza, J.G. Soares, M.F. Da Silva, Thermal and photochemicalmethods for the preparation of thin films of cermet materials, J. Mater. Sci. 31(1996) 2325–2332.

[46] V. Subramanian, E. Wolf, P.V. Kamat, Catalysis with TiO2/gold nanocompo-sites. Effect of metal particle size on the fermi level equilibration, J. Am.Chem. Soc. 126 (2004) 4943–4950.

[47] J.S. Curran, D. Lamouche, Transport and kinetics in photoelectrolysis bysemiconductor particles in suspension, J. Phys. Chem. 87 (1983) 5405–5411.

[48] V. Subramanian, E. Wolf, P.V. Kamat, Green emission to probe photoinducedcharging events in ZnO�Au Nanoparticles. Charge distribution and fermi-level equilibration, J. Phys. Chem. B 107 (2003) 7479–7485.

[49] A. Wood, M. Giersig, P. Mulvaney, Fermi level equilibration in quantumdot�metal nanojunctions, J. Phys. Chem. B 105 (2001) 8810–8815.

[50] M. Jakob, H. Levanon, Charge distribution between UV-irradiated TiO2 andgold nanoparticles: determination of shift in the fermi level, Nano. Lett. 3(2003) 353–358.

[51] S. Anandan, P. Sathish Kumar, N. Pugazhenthiran, J. Madhavan,P. Maruthamuthu, Effect of loaded silver nanoparticles on TiO2 forphotocatalytic degradation of Acid Red 88, Sol. Energy Mater. Sol. Cells92 (2008) 929–937.

[52] J. Flanagan, W.P. Griffith, A.C. Skapski, The active principle of Caro’s acid,HSO5

–: X-ray crystal structure of KHSO5 �H2O, J. Chem. Soc. Chem. Commun.23 (1984) 1574–1575.

[53] B. Meunier, Potassium monopersulfate: just another primary oxidant or ahighly versatile oxygen atom donor in metalloporphyrin-mediated oxygena-tion and oxidation reactions, New J. Chem. 16 (1992) 203–211.

[54] M.A. Fox, M.T. Dulay, Heterogeneous photocatalysis, Chem. Rev. 93 (1993)341–357.