role of phosphorus in synthesis of phosphated mesoporous tio2 photocatalytic materials by eisa...

8
Role of phosphorus in synthesis of phosphated mesoporous TiO 2 photocatalytic materials by EISA method Xiaoxing Fan a,b , Tao Yu a,c , Ying Wang a,d , Jing Zheng a,b , Ling Gao d , Zhaosheng Li a,b,c , Jinhua Ye e , Zhigang Zou a,c, * a Eco-materials and Renewable Energy Research Center (ERERC), Department of Physics, Nanjing University, Nanjing 210093, People’s Republic of China b Department of Materials Science and Engineering, Nanjing University, Nanjing 210093, People’s Republic of China c National Laboratory of Solid State Microstructures, Nanjing University, Nanjing 210093, People’s Republic of China d School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, People’s Republic of China e Photocatalytic Materials Center (PCMC), National Institute for Materials Science (NIMS), 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047, Japan Received 5 July 2007; received in revised form 23 January 2008; accepted 9 February 2008 Available online 20 February 2008 Abstract The phosphated mesoporous TiO 2 (PMT) were synthesized by using evaporation-induced self-assembly approach (EISA) with phosphorus content from 1 to 15 mol%. The X-ray diffraction and N 2 adsorption–desorption isothermal results reveal that the incorporating of phosphorus is of benefit to improving the thermal stability and enhancing the surface area of mesoporous TiO 2 by constraining the growth of anatase crystallite. XPS confirms the phosphorus in the calcined PMT exists as amorphous titanium phosphate in a pentavalent-oxidation state (P 5+ ) and embedded into the nanocrystalline anatase TiO 2 . In photodegradation gas phase acetaldehyde, the photocatalytic activity of PMT samples is higher than that of pure mesoporous TiO 2 and P25. It is believed that the enhancing photocatalytic activity of phosphated mesoporous TiO 2 is mainly caused by two factors relative with the incorporating of phosphorus in framework. # 2008 Elsevier B.V. All rights reserved. Keywords: Phosphated mesoporous TiO 2 ; Amorphous titanium phosphate; Photocatalysis; Acetaldehyde 1. Introduction In the past three decades, much attention has been paid to the metal-oxide photocatalysts due to their possible application in degradation of environmental pollutants [1–4] and conversion of solar energy [5,6]. Among all these photocatalysts, P25 (TiO 2 ) is one of the most popular and promising presently known material because of its high photocatalytic activity, non- toxicity, long-term stability, and low price [7]. Recently, mesoporous TiO 2 has attracted much interest because it performs better photocatalytic properties than P25 [8–12]. The higher photocatalytic activity of mesoporous TiO 2 can be attributed to the mesoporous structure that offers more active sites for catalytic reactions and improves diffusion of gaseous molecules [12,13]. Unfortunately, it is difficult to synthesize a mesoporous TiO 2 possessing both large surface area and high crystallite. Since during the process of removing surfactants, solvent extraction or calcination at lower temperature usually leads to an inactive amorphous TiO 2 [14,15]. In order to remove the surfactants completely and obtain high-crystalline meso- porous TiO 2 , the higher calcination temperatures are always adopted. But high temperature may cause the collapse of the mesoporous framework and decrease the surface area. Much effort, for example, incorporating phosphorus, has been devoted to increase the surface area and improve the thermal stability of the mesoporous TiO 2 materials. In 1998, Stone and Davies synthesized large surface area but poor crystalline mesoporous TiO 2 by using dodecyl phosphate surfactant. The poor crystallinity is due to the remained phos- phorus from surfactant that very strongly bond to the titania wall [14]. The photocatalytic activity of this amorphous photocatalyst in the photo-oxidation of 2-propanol to acetone was much lower than that on P25, the low activity was attributed to the detrimental www.elsevier.com/locate/apsusc Available online at www.sciencedirect.com Applied Surface Science 254 (2008) 5191–5198 * Corresponding author at: Eco-materials and Renewable Energy Research Center (ERERC), Department of Physics, Nanjing University, Nanjing 210093, People’s Republic of China. Fax: +86 25 8368 6632. E-mail address: [email protected] (Z. Zou). 0169-4332/$ – see front matter # 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2008.02.038

Upload: independent

Post on 14-May-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

www.elsevier.com/locate/apsusc

Available online at www.sciencedirect.com

4 (2008) 5191–5198

Applied Surface Science 25

Role of phosphorus in synthesis of phosphated mesoporous TiO2

photocatalytic materials by EISA method

Xiaoxing Fan a,b, Tao Yu a,c, Ying Wang a,d, Jing Zheng a,b, Ling Gao d,Zhaosheng Li a,b,c, Jinhua Ye e, Zhigang Zou a,c,*

a Eco-materials and Renewable Energy Research Center (ERERC), Department of Physics, Nanjing University, Nanjing 210093,

People’s Republic of Chinab Department of Materials Science and Engineering, Nanjing University, Nanjing 210093, People’s Republic of Chinac National Laboratory of Solid State Microstructures, Nanjing University, Nanjing 210093, People’s Republic of China

d School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, People’s Republic of Chinae Photocatalytic Materials Center (PCMC), National Institute for Materials Science (NIMS), 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047, Japan

Received 5 July 2007; received in revised form 23 January 2008; accepted 9 February 2008

Available online 20 February 2008

Abstract

The phosphated mesoporous TiO2 (PMT) were synthesized by using evaporation-induced self-assembly approach (EISA) with phosphorus

content from 1 to 15 mol%. The X-ray diffraction and N2 adsorption–desorption isothermal results reveal that the incorporating of phosphorus is of

benefit to improving the thermal stability and enhancing the surface area of mesoporous TiO2 by constraining the growth of anatase crystallite. XPS

confirms the phosphorus in the calcined PMT exists as amorphous titanium phosphate in a pentavalent-oxidation state (P5+) and embedded into the

nanocrystalline anatase TiO2. In photodegradation gas phase acetaldehyde, the photocatalytic activity of PMT samples is higher than that of pure

mesoporous TiO2 and P25. It is believed that the enhancing photocatalytic activity of phosphated mesoporous TiO2 is mainly caused by two factors

relative with the incorporating of phosphorus in framework.

# 2008 Elsevier B.V. All rights reserved.

Keywords: Phosphated mesoporous TiO2; Amorphous titanium phosphate; Photocatalysis; Acetaldehyde

1. Introduction

In the past three decades, much attention has been paid to the

metal-oxide photocatalysts due to their possible application in

degradation of environmental pollutants [1–4] and conversion

of solar energy [5,6]. Among all these photocatalysts, P25

(TiO2) is one of the most popular and promising presently

known material because of its high photocatalytic activity, non-

toxicity, long-term stability, and low price [7]. Recently,

mesoporous TiO2 has attracted much interest because it

performs better photocatalytic properties than P25 [8–12]. The

higher photocatalytic activity of mesoporous TiO2 can be

attributed to the mesoporous structure that offers more active

sites for catalytic reactions and improves diffusion of gaseous

* Corresponding author at: Eco-materials and Renewable Energy Research

Center (ERERC), Department of Physics, Nanjing University, Nanjing 210093,

People’s Republic of China. Fax: +86 25 8368 6632.

E-mail address: [email protected] (Z. Zou).

0169-4332/$ – see front matter # 2008 Elsevier B.V. All rights reserved.

doi:10.1016/j.apsusc.2008.02.038

molecules [12,13]. Unfortunately, it is difficult to synthesize a

mesoporous TiO2 possessing both large surface area and high

crystallite. Since during the process of removing surfactants,

solvent extraction or calcination at lower temperature usually

leads to an inactive amorphous TiO2 [14,15]. In order to remove

the surfactants completely and obtain high-crystalline meso-

porous TiO2, the higher calcination temperatures are always

adopted. But high temperature may cause the collapse of the

mesoporous framework and decrease the surface area. Much

effort, for example, incorporating phosphorus, has been

devoted to increase the surface area and improve the thermal

stability of the mesoporous TiO2 materials.

In 1998, Stone and Davies synthesized large surface area but

poor crystalline mesoporous TiO2 by using dodecyl phosphate

surfactant. The poor crystallinity is due to the remained phos-

phorus from surfactant that very strongly bond to the titania wall

[14]. The photocatalytic activity of this amorphous photocatalyst

in the photo-oxidation of 2-propanol to acetone was much lower

than that on P25, the low activity was attributed to the detrimental

X. Fan et al. / Applied Surface Science 254 (2008) 5191–51985192

effect of surface defects on amorphous photocatalyst. After-

wards, Yu et al. synthesized phosphated mesoporous titanium

dioxide by incorporating phosphorus from H3PO4 directly into

the framework of TiO2. They found that the incorporation of

phosphorus could stabilize the TiO2 framework of mesoporous

TiO2 when calcined at high temperature. In their study, the

phosphated mesoporous TiO2 exhibited a higher photocatalytic

activity than P25 [10]. Recently, Dekany et al. prepared a

phosphated mesoporous TiO2 with high surface area by varying

the P/Ti molar ratio between 1 and 10 mol% in a non-template

sol–gel system. An optimum phosphorus content corresponding

to high photocatalytic activity in phosphated mesoporous TiO2

was obtained [16]. Surprisingly, very little attention has been

paid to clarify the relationship of phosphorus content with the

surface area and give a conclusion on the mechanism of

enhancing photocatalytic activity of the phosphated mesoporous

TiO2. Thus, it is necessary to judge the delicate role of

phosphorus in mesoporous TiO2 and obtain further insight into

material synthesize science and photocatalysis.

In the current work, phosphated mesoporous TiO2 (PMT)

photocatalyst were synthesized by an evaporation-induced self-

assembly approach (EISA) which is generally adept to synthesize

mesoporous transition metal oxide [17,18]. To investigate the

effects of phosphorous content on the surface area of meso-

structure, the content of phosphorus was changed in a wide range

from 1 to 15 mol%. High phosphorus content PMT (100 mol% of

P/Ti molar ratio) was also synthesized for deducting the

mechanism of enhancement in photocatalytic activity of PMT.

The photocatalytic activities of these samples were valued on

acetaldehyde photodegradation under UV irradiation.

2. Experimental

2.1. Preparation of catalysts

The samples of PMT were prepared by an EISA process. In a

typical synthesis, 1 g of pluronic P123 (EO20PO70EO20,

M = 5800, Aldrich) was dissolved in 10 g of ethanol, then

0.01 mol of titanium chloride (TiCl4) was added with vigorous

stirring for 0.5 h. After then, 1 � 10�4, 5 � 10�4, 1 � 10�3 and

1.5 � 10�3 mol H3PO4 was added to above solution for

synthesizing 1 mol% (0.84 wt%, P2O5wt%), 5 mol%

(4.2 wt.%), 10 mol% (8.4 wt.%) and 15 mol% (12.7 wt.%)

phosphated samples. The resulting sol was gelled in an open

Petri dish at 40 8C oven for 4 days (static state). The as-made bulk

sample was then calcined at 400, 600 and 800 8C for 5 h in air

with the heating rate of 3 8C min�1 to remove the surfactant,

respectively. The calcined samples were labeled according to its

phosphorus content and calcination temperature. For example,

pure mesoporous TiO2 calcined at 400 8C was labeled as MT-0-

400. Similarly, the label of MT-1-400 indicates the sample

containing 1 mol% phosphorus and calcined at 400 8C.

2.2. Structure characterization

X-ray powder diffraction measurements were performed on

a Rigaku Ultima III X-ray diffractometer using Cu Ka radiation

(l = 1.54056 A). Nitrogen adsorption–desorption isotherms

were collected on a Micromeritics Tristar-3000 surface area

and porosity analyzer at 77 K after the samples had been

degassed in the flows of N2 at 180 8C for 5 h. The Brunauer–

Emmett–Teller (BET) surface area was calculated from the

linear part of the BET plot (P/P0 = 0.1–0.25). The pore size

distribution plots were obtained by using the Barret–Joyner–

Halenda (BJH) model. Images of high-resolution transmission

electron microscope (HRTEM) were obtained by employing a

FEI Tecnai G2 20 S-TWIN high-resolution transmission

electron microscope with a 200 kV accelerating voltage. The

samples for TEM were prepared by dispersing the final powders

in ethanol and the dispersion was dropped on carbon-copper

grids. The X-ray photoelectron spectroscopy (XPS) measure-

ments were carried out on an ESCALAB Mark II (VG

Company, UK); all binding energies were referenced to the C1s

peak at 284.6 eV of the surface adventitious carbon. The

infrared spectroscopy spectra were obtained on Brucker Vector

22. The UV–vis diffuse reflectance spectrum was recorded with

a UV–vis spectrometer (UV-2550, Shimadzu) at room

temperature and transformed to the absorption spectrum

according to the Kubelka–Munk relationship. The absorption

band edge was determined by extrapolating the linear part of

the plot to wavelength axis (Abs. = 0).

2.3. Photocatalytic activities evaluation

The photocatalytic activities of the calcined samples were

performed by oxidizing acetaldehyde in air using a gastight

system with a quartz window. In a typical process, 0.1 g of

powder photocatalysts with adding several drops of water was

smeared on a 4-cm2 glass groove and dried in oven. The

purpose of adding little water is to forming a uniform coating

adhering on glass substrate. The glass with powder photo-

catalyst was then put into a 224-ml reactor, filled with air to one

atmospheric pressure. Then, acetaldehyde (10 ml of 40%

CH3CHO aqueous solution) was injected into the reactor to

generate a high-concentration acetaldehyde gas. The light

source for the catalytic reaction was utilized a 300-W Xe arc

lamp. In addition, a gas pump was used for accelerating gas

diffusion. Carbon dioxide was detected by a gas chromato-

graphy (CO2, GC-8A with TCD detector, Shimadzu).

3. Results and discussion

3.1. Powder X-ray diffraction analysis

Fig. 1 shows the wide-angle XRD patterns of PMT and MT

calcined at different temperature. When calcined at 800 8C, two

peaks at 22.558 (d = 3.9 A) and 27.658 (d = 3.2 A) appeared in

spectrum of MT-15-800 correspond to the (600) and (721) XRD

diffraction peak of crystalline TiP2O7 (JCPDS, No. 38-1468).

The intensity of the two peaks was weak in XRD pattern of MT-

10-800 (see Supplementary data). The appearance of TiP2O7

peaks in MT-10-800 and MT-15-800 indicates that the

phosphorus still remained in PMT samples and did not

evaporate even at 800 8C calcination. At low calcination

Fig. 1. Wide-angle XRD patterns of pure mesoporous TiO2 and PMT of

containing 1 mol% phosphorus calcined at 400, 600 and 800 8C and MT-15-

800.Fig. 2. N2 adsorption–desorption isotherms and Barret–Joyner–Halenda (BJH)

pore size (inset) of pure mesoporus TiO2 and 1 mol% phosphorus content of

PMT.

X. Fan et al. / Applied Surface Science 254 (2008) 5191–5198 5193

temperature of 400 and 600 8C, only pure anatase phases

(JCPDS, No. 21-1272) and no obvious XRD peaks corre-

sponded to titanium phosphate are observed for PMT samples.

The absence of TiP2O7 peaks in sample calcined at 400 and

600 8C may indicate the phosphorus exists as amorphous

phosphate. Additionally, the rutile (2u = 27.458) phase emerged

in MT, but absent in all spectra of PMT samples [19]. Table 1

summarizes the crystalline size of the samples calcined at

different temperatures, which is calculated from the Scherrer

equation by using the (1 0 1) XRD peaks of anatase phase. At

calcination temperature of 400 8C, the crystallite size of MT-0-

400 is 12.4 nm. However, those of the PMT samples are not

larger than 7.9 nm. All wide-angle XRD patterns show that the

size of crystallite becomes larger with the increase of the

calcination temperature. The results of XRD characterization

obviously clarify that the function of phosphorus in framework

is not only constraining the form of rutile phase but also

inhibiting the growth of anatase TiO2 crystal. It is well known

that the growth of nanocrystal usually leads to the collapse of

Table 1

Summary of the physicochemical Properties of samples calcined at different temp

Samples 400 600

Surface area

(m2 g�1)

Pore

size (nm)

Crystalline

size (A)

Surface

area (m2 g�1)

MT-0- 111 8.4 124 (4) 52

MT-1- 171 6.8 79 (1) 106

MT-5- 206 6.3 72 (2) 134

MT-10- 217 6.5 75 (2) 142

MT-15- 208 5.1 68 (4) 125

mesoporous framework, which broadens the diameter distribu-

tion of channels [20]. In other words, phosphorus in framework

can increase the thermal stability and surface area by inhibiting

the growth of grains and protect the mesoporous structure from

collapsing.

3.2. Nitrogen physisorption

The pore size distributions and N2 adsorption–desorption

isotherms of PMT calcined at different temperatures are shown

in Fig. 2. All the isotherms are type IV with a clear

characteristic hysteresis loop of mesoporous materials. The

BET surface areas of samples calcined at different temperatures

are summarized in Table 1. At the same calcination temperature

of 400 8C, the surface area of MT-1-400, MT-5-400, MT-10-

400, MT-15-400 is 171, 206, 217 and 208 m2 g�1, respectively,

which is much larger than that of MT-0-400 (111 m2 g�1). The

eratures

800

Pore

size (nm)

Crystalline

size (A)

Surface

area (m2 g�1)

Pore

size (nm)

Crystalline

size (A)

8.8 224 (7) 7 21.6 592 (14)

8.7 117 (2) 41 18.5 221 (3)

7.2 88 (2) 50 15.9 172 (2)

6.8 86 (2) 34 19.0 177 (2)

7.0 90 (3) 22 19.0 192 (2)

X. Fan et al. / Applied Surface Science 254 (2008) 5191–51985194

surface area of MT-1-400 increases by 54% through

incorporated 1 mol% P, and increases further by 20%

incorporated 5 mol% P, and do not increase further when the

content of P is over 5 mol%. At higher calcined temperature of

600 and 800 8C, the PMT samples also characterized a larger

surface area than that of mesoporous TiO2. Based on the results

of N2 adsorption isothermal, the relation between the BET

surface areas and phosphate content could be clarified. Take

example for the PMT samples calcined at 400 8C, the surface

area increased monotonously with phosphate content (phos-

phorus content<5 mol%), then slowly increased when the ratio

of phosphor/titanium over 5 mol%. The large surface area of

PMT samples were attributed to their small crystalline size

caused by phosphorus incorporation. The surface area of MT-

15-400 is slight smaller than MT-10-400 is probably due to the

remaining carbon caused by more phosphorus incorporation.

3.3. TEM and HRTEM characterization

TEM and HRTEM observation was performed. In the TEM

images (Fig. 3), a wormhole-like mesostructure without long-

range order was observed in MT-0-400, MT-1-400, MT-5-400

and MT-1-600. In contrast with MT-0-400, the sample of MT-1-

400 and MT-5-400 show a thinner pore wall which results in a

high surface area. Because of the growth of grain, the pore wall

of MT-1-600 is thicker than MT-1-400, indicating that a certain

extent collapses of mesostructure at high calcination tempera-

Fig. 3. TEM images of MT-0-400 (a), MT-1-

ture. As shown in HRTEM images (Fig. 4), the nanocrystal

anatase TiO2 with a clear lattice (d = 3.5 A) connect with each

other to form crystalline framework walls of the mesopore. It

was estimated that the size of nanocrystals of MT-0-400, MT-1-

400, MT-5-400 and MT-1-600 is approximately 13, 8, 8 and

14 nm as marketed in Fig. 4, respectively. This is in agreement

with the result of XRD. Additionally, no crystalline phosphate

lattice was observed on these PMT samples and only anatase

lattice was clear shown in HRTEM, it is speculated that the

amorphous titanium phosphate should embedded in the TiO2

crystalline grains and acted as a diffusion barrier to inhibit

crystal growth of TiO2 during calcinations.

3.4. Infrared spectroscopy (FT-IR) and XPS performance

The FT-IR spectra of MT-400, MT-1-400, MT-5-400 and

MT-10-400 are shown in Fig. 5. The bond at 1600 cm�1

appeared in the spectra of all samples is corresponded to O–H

stretching vibration due to the surface-absorbed water and

hydroxyl groups [21,22], while the bond at 1100 cm�1 is

assigned to Ti–O–P vibrations that corresponds to the

phosphate in the frame of TiO2 [9,23,24]. The intensity of

absorption peak (1100 cm�1) turns stronger may be result form

the increasing concentration of phosphate content in mesos-

tructure. As expected, the bond at 1100 cm�1 was absent in

pure mesoporous TiO2. It was also noticed that the

transmittance of all curves were gradually weaken with the

400 (b), MT-5-400 (c) and MT-1-600 (d).

Fig. 4. HRTEM images of MT-0-400 (a), MT-1-400 (b), MT-5-400 (c) and MT-1-600 (d).

X. Fan et al. / Applied Surface Science 254 (2008) 5191–5198 5195

increasing phosphorus content in PMT samples when

wavenumber was lower than 900 cm�1 corresponding to Ti–

O–Ti (ca. 850 cm�1). This can be explained that the relative

content of pure TiO2 is decreased result from the increasing

concentration of amorphous phosphated titania. Fig. 6 shows

the full XPS spectra of MT-10-400, which indicates the PMT

sample contains only Ti, O, C and P. The high-resolution XPS

spectra of Ti2p show that the peak of Ti2p (458.05 eV) in MT

samples shifts toward higher energy (458.15 eV) in PMT,

Fig. 5. FT-IR spectrum of MT-0-400 (a), MT-1-400 (b), MT-5-400 (c) and MT-

10-400 (d).

which indicates that the local chemical environment of Ti has

been significantly influenced by phosphorus incorporation in

amorphous titanium phosphate. While the binding energy of P2p

peak centered at 133 eV suggests that the phosphorus in calcined

PMT exists in a pentavalent-oxidation state (P5+). Since H3PO4

was used as the starting material to synthesize phosphated TiO2.

There are three possible chemical states of amorphous titanium

phosphate, including Ti3(PO4)4, Ti(HPO4)2 and Ti(H2PO4)4. The

high-resolution XPS spectrum of P2p reveals the derivation of

three peaks at 132.4, 133.4, and 134.7 eV, which correspond

to the binding energies of PO43�, HPO4

2�, and H2PO4�,

respectively. According to the area of three peaks, the phosphorus

atomic ratio of PO43�, HPO4

2�, and H2PO4� is calculated to be

12:76:12, which implies that the main phosphoric type presented

in the phosphated mesoporous TiO2 is HPO42� [10]. On the basis

of the results of XRD, phosphorus exists as amorphous titanium

phosphate that embedded in the TiO2 crystalline grain at low

calcination temperature (<600 8C), while the FT-IR spectra

support a Ti–O–P structure in the titanium phosphate samples.

From the information given by XPS and IR spectra, It is

speculated that amorphous titanium phosphate is mainly

composed of Ti(HPO4)2.

3.5. Photocatalytic activity studies

Fig. 7 shows the UV–vis absorption spectra of phosphated

mesoporous TiO2 calcined at 400 8C. The absorption band

edge of phosphorus free sample MT-0-400 is 379 nm, while

Fig. 6. The full XPS, high-resolution Ti2p and P2p XPS spectra of MT-10-400.

X. Fan et al. / Applied Surface Science 254 (2008) 5191–51985196

almost two overlapped absorption spectra with a band edge of

376 nm were obtained on MT-1-400 and MT-5-400. Mean-

while no large band edge (at 377 nm) change was found on the

spectrum of MT-10-400 even the phosphorous content up to

10%. A small red shift to 384 nm was observed on the MT-15-

400, which may be caused by carbon remaining. These UV–

vis absorption spectra reveal that the phosphorus does not

exist as dopant in lattice of crystalline TiO2 since phosphorus

doping into lattice of TiO2 usually leads to a visible light

absorption up to 450 nm due to the formation of impurity

level [25].

Acetaldehyde, as a well-known indoor air pollutant, is

largely formed as an intermediate during photocatalytic

oxidation of other organic compounds. The photocatalytic

properties of PMT were investigated under Xe lamp irradiation

without filter (full arc irradiation). The spectrum of Xe lamp is

shown in Fig. 7 (inset).

Fig. 8 displays the increasing rate of carbon dioxide

produced from the photocatalytic degradation of acetaldehyde

on MT-1-400, MT-5-400, MT-10-400, MT-15-400 and MT-0-

Fig. 7. UV–vis diffuse reflectance spectra of mesoporous TiO2 samples cal-

cined at 400 8C. MT-0-400 (a), the superposition curve of MT-1-400 and MT-5-

400 (b), MT-10-400 (c), MT-15-400 (d), the inset is the spectrum of Xe lamp.

400 under UV light irradiation. For comparison, 0.1 g of SiO2

was placed into photocatalytic reactor and tested under the

same conditions as that of on PMT samples, no increase of CO2

concentration was observed. This proves that the produced CO2

comes from the photodecomposition of acetaldehyde on PMT

samples. The CO2 product comes from the acetaldehyde

oxidation in photocatalytic reactor can be depicted in Eq. (1).

2CH3CHOþ 5O2 �!TiO2;hn

4CO2 þ 4H2O (1)

As expected, the photocatalytic activity of PMT to oxidize

acetaldehyde under UV irradiation was affected by the

amorphous titanium phosphate. In 25 min photocatalytic

reaction, the formation rates of carbon dioxide are 5.78,

4.62, 4.30, 4.22, 4.06 and 2.98 mmol min�1 for MT-1-400, MT-

5-400, MT-10-400, MT-15-400, MT-0-400 and P25, respec-

tively (see Fig. 8). The photocatalytic conversion ratios of

Fig. 8. The photocatalytic activities of P25, PMT and pure mesoporous TiO2

calcined at 400 8C.

Scheme 1. Characterization of phosphated mesoporous titanium by using P123

surfactant H3PO4 as phosphorus source.

X. Fan et al. / Applied Surface Science 254 (2008) 5191–5198 5197

acetaldehyde to carbon dioxide were 82, 66, 61, 60, 58 and 42%

for MT-1-400, MT-5-400, MT-10-400, MT-15-400, MT-0-400

and P25, respectively. The activities of PMT are gradually

decreasing with the increase of titanium phosphate content in

mesoporous TiO2. The photocatalytic activity of MT-1-400 is

40% higher than that of MT-0-400 and 90% higher than that of

P25. In the same photocatalytic reaction condition, the

photocatalytic activity of amorphous titanium phosphate is

only 0.16 mmol min�1 even though it possesses a higher BET

surface area than 200 m2/g. It is noticed that the affection of

incorporating phosphorus into mesoporous TiO2 in our PMT

samples is different from that reported by Stone and Davis [14].

Their investigation revealed that the photocatalytic activity of

PMT synthesized by using dodecyl phosphate surfactant as a

template was much lower than that of P25. While similarly to

our results, an improve affection of incorporating phosphorus

on the photocatalytic activity of PMT synthesized by directly

using H3PO4 as phosphorus source was reported [10,16]. These

contradictory conclusions imply that phosphorus acts as a

complex role in synthesis process. Different phosphorous

precursor will influence the surface properties and crystalline of

the resulted photocatalysts. As well known, photocatalytic

reaction is carried out on the surface of photocatalyst. The

larger the surface area of the photocatalyst is, the more reaction

sites are, which is in favor of the activity. Crystallinity is

another important factor to influence the ability because the

crystal defects are always recombination centers of photo-

generated electrons and holes. From the synthesis process

reported by Stone and Davis, it is speculated that the surface of

PMT synthesized by using phosphate surfactant was mainly

coated with amorphous titanium phosphate. Since the forma-

tion of mesoporous materials was directed by surfactants, the

anionic titaniums were deposited around the positively charged

surfactant templated to form inorganic–organic mesostructures

via electrostatic interactions. When using phosphate surfac-

tants, after calcination, the phosphorus will react with the

titanium on the surface of pore wall and form amorphous

titanium phosphate. While when directly incorporating H3PO4

and using non-phosphate surfactant such as P123, a different

structure was obtained. In this synthesis condition, the

phosphorus was dispersed in the framework of mesoporous

TiO2 and exists as amorphous titanium phosphate, which

mainly embedded in the TiO2 crystalline grains.

It was known that, amorphous materials, as well as amorphous

titanium phosphate, generally contain many imperfections, i.e.,

impurities, dangling bonds, or microvoids, which lead to

electronic states in the band gap, and they act as a recombination

center of e� and h+ [15]. Therefore, the existence of amorphous

titanium phosphate is disadvantage to the photocatalytic activity

by decreasing the quantity of e� and h+ for photocatalytic

reaction. Basis on the analysis above, it is concluded that the low

photocatalytic activity of PMT synthesized by using phosphate

surfactant may be resulted from its surface that was full coated by

amorphous titanium phosphate which inhibited the inside e�

and h+ transporting to surface, thus lead to performing low

photocatalytic activity. Additionally, the low photocatalytic

activity is also partly caused by the amorphous titanium

phosphate which cannot provide photogenerated e� and h+. On

the contrary, the PMT samples by directly using H3PO4

performed high photocatalytic activity. This is due to the

different position of amorphous phosphate. As illuminated in

Scheme 1, the amorphous phosphate was embedded into the

nanocrystalline anatase TiO2 but not on the surface completely.

Thus, the inside e� and h+ can easily transport to surface through

crystalline TiO2.

Contrasted with pure mesoporous TiO2, as shown in Fig. 7,

there is no difference in photoabsorption for mesoporous TiO2

and phosphated mesoporous TiO2 calcined at 400 8C. The

photocatalytic activity of PMT containing 1 mol% phosphorus is

about two times higher than that of P25 in photodegradation gas

phase acetaldehyde. The higher activity can be attributed to the

balance of two factors. One is the large surface area caused by

incorporating phosphorus in framework, the large surface area

provided more active sites for catalysis and enhanced the

photocatalytic activity, thus, the phosphorus play a positive role.

The other is the concentration of amorphous titanium phosphate

in framework of mesostructure. The amorphous titanium

phosphate is a poor photocatalyst itself; it does not provide

much e� and h+ when irradiated by ultraviolet radiation. In

addition, the amorphous titanium phosphate acts as a recombi-

nation centers by trapping photogenerated electrons e� or holes

h+ and inhibits the inside e� and h+ transport to surface.

Therefore, incorporating phosphorus will be of benefit to

improving photocatalytic activity of mesoporous when the two

the factors balanced at optimum phosphorus content. On the

basis of our photocatalysis performs, the content of incorporated

1 mol% P may be very close to the optimum.

4. Conclusions

Phosphated mesoporous TiO2 materials have been synthe-

sized by using evaporation-induced self-assembly method. The

X. Fan et al. / Applied Surface Science 254 (2008) 5191–51985198

incorporation of phosphorus in mesoporous TiO2 constrains the

growth of crystals of PMT and improves thermal stability. The

surface areas increased remarkably with the increase of phos-

phorus content when phosphorus content was less than 5 mol%.

Over 10 mol% phosphorus incorporating, the function is not

obviously for enhancing surface area and thermal stability of

PMT. The increasing surface area of PMT is caused by the incor-

porated phosphorus inhibiting the anatase growth. The phos-

phorus species exist as amorphous titanium phosphate which

embed in anatase nanocrystal and is mainly composed of

Ti(HPO4)2.

The incorporating phosphorus exists as amorphous titanium

play two roles. One is enhancing the surface area, which is

beneficial for photocatalytic activity. The other is forming

recombination center of e� and h+ and decreasing photocatalytic

activity. The optimum phosphorus content of 1 mol% P/Ti mole

ratio is the balance of two factors for photocatalytic activity.

Acknowledgements

Financial support from the National Natural Science

Foundation of China (Nos. 20603017 and 20528302), the

National High Technology Research and Development Program

of China (No. 2006AA05Z113), the Science and Technology

Research Program of the Ministry of Education (MOE) of China

(No. 307012) and the National Basic Research Program of China

(973 Program, 2007CB613301, 2007CB613305) is gratefully

acknowledged. This work was also supported by the Scientific

Research Foundation of Graduate School of Nanjing University

(2006CL05). Prof. Z.G. Zou and T. Yu would like to thank the

Jiangsu Provincial Talent Scholars Program.

Appendix A. Supplementary data

Supplementary data associated with this article can be found,

in the online version, at doi:10.1016/j.apsusc.2008.02.038.

References

[1] A. Fujishima, K. Honda, Nature 238 (1972) 37–38.

[2] A. Mills, S.L. Hunte, J. Photochem. Photobiol. A 108 (1997) 1–35.

[3] D.A. Tryk, A. Fujishima, K. Honda, Electrochim. Acta 45 (2000) 2363–

2376.

[4] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95

(1995) 69–96.

[5] Z. Zou, J. Ye, K. Sayama, H. Arakawa, Nature 414 (2001) 625–627.

[6] A. Hagfeldt, M. Gratzel, Acc. Chem. Res. 33 (2000) 269–277.

[7] T. Kawahara, Y. Konishi, H. Tada, N. Tohge, J. Nishii, S. Ito, Angew.

Chem. Int. Ed. 41 (2002) 2811–2813.

[8] J.C. Yu, L. Zhang, Z. Zheng, J. Zhao, Chem. Mater. 15 (2003) 2280–

2286.

[9] J.C. Yu, L. Zhang, J. Yu, Chem. Mater. 14 (2002) 4647–4653.

[10] L. Saadoun, J.A. Ayllon, J. Jimenez-Becerril, J. Peral, X. Domenech, R.

Rodrıguez-Clemente, Mater. Res. Bull. 35 (2000) 193–202.

[11] X. Wang, J.C. Yu, C. Ho, Y. Hou, X. Fu, Langmuir 21 (2005) 2552–

2559.

[12] D.M. Antonelli, J.Y. Ying, Angew. Chem. Int. Ed. Engl. 34 (1995) 2014–

2017.

[13] J.Y. Ying, AIChE J. 46 (2000) 1902–1906.

[14] V.F. Stone Jr., R.J. Davis, Chem. Mater. 10 (1998) 1468–1474.

[15] B. Ohtani, Y. Ogawa, S.I. Nishimoto, J. Phys. Chem. B 101 (1997) 3746–

3750.

[16] L. Korosi, I. Dekany, Colloids Surf. A: Physicochem. Eng. Aspects 280

(2006) 146–154.

[17] G.J. de, A.A. Soler-Illia, A. Louis, C. Sanchez, Chem. Mater. 14 (2002)

750–759.

[18] P. Yang, D. Zhao, D.I. Margolese, B.F. Chmelka, G.D. Stucky, Nature 396

(1998) 152–155.

[19] E. Ortiz-Islas, R. Gomez, T. Lopez, J. Navarrete, D.H. Aguilar, P.

Quintana, Appl. Surf. Sci. 252 (2005) 807–812.

[20] B. Smarsly, D. Grosso, T. Brezesinski, N. Pinna, C. Boissiere, M.

Antonietti, C. Sanchez, Chem. Mater. 16 (2004) 2948–2952.

[21] M.P. Kapoor, S. Inagaki, H. Yoshida, J. Phys. Chem. B 109 (2005) 9231–

9238.

[22] Z. Ding, G.Q. Lu, P.F. Greenfield, J. Phys. Chem. B 104 (2000) 4815–

4820.

[23] A. Bhaumik, S. Inagaki, J. Am. Chem. Soc. 123 (2001) 691–696.

[24] Q. Wang, L. Zhong, J. Sun, J. Shen, Chem. Mater. 17 (2005) 3563–3569.

[25] Q. Shi, D. Yang, Z. Jiang, J. Li, J. Mol. Catal. B: Enzyme 43 (2006) 44–48.