size dependency of nanocrystalline tio2 on its optical property and

11
Size dependency of nanocrystalline TiO 2 on its optical property and photocatalytic reactivity exemplified by 2-chlorophenol H. Lin a , C.P. Huang a, * , W. Li b , C. Ni b , S. Ismat Shah b,c , Yao-Hsuan Tseng d a Department of Civil and Environmental Engineering, University of Delaware, Newark, DE 19716, United States b Department of Materials Sciences and Engineering, University of Delaware, Newark, DE 19716, United States c Department of Physics and Astronomy, University of Delaware, Newark, DE 19716, United States d Energy and Environment Laboratories, Industrial Technology Research Institute, Hsinchu 301, Taiwan, ROC Received 19 January 2006; received in revised form 7 July 2006; accepted 18 July 2006 Available online 1 September 2006 Abstract Anatase TiO 2 nanocrystallines (17–29 nm) were successfully synthesized by the metal–organic chemical vapor deposition method (MOCVD). Moderate manipulation of system parameters of MOCVD can control the particle size. The electro-optical and photocatalytic properties of the synthesized TiO 2 nanoparticles were studied along with several commercially available ultra-fine TiO 2 particles (e.g., 3.8–5.7 nm). The band gap of the TiO 2 crystallines was determined using the transformed diffuse reflectance technique according to the Kubelka–Munk theory. Results showed that the band gap of TiO 2 monotonically decreased from 3.239 to 3.173 eV when the particle size decreased from 29 to 17 nm and then increased from 3.173 to 3.289 eV as the particle size decreased from 17 to 3.8 nm. The results of band gap change as a function of particle size agreed well with what was predicted by the Brus’ equation, i.e., the effective mass model (EMM). However, results of the photocatalytic oxidation of 2-chlorophenol (2-CP), showed that the smaller the particle size, the faster the degradation rate. This is attributed in part to the combined effect of band gap change relative to the spectrum of the light source and the specific surface area (or particle size) of the photocatalysts. The change of band gap due to particle size represents only a small optical absorption window with respect to the total spectrum of the light source, i.e., from 380 to 400 nm versus >280 nm. Consequently, the gain in optical property of the larger particles was severely compromised by their decrease in specific surface area. Our results clearly indicated the importance of specific surface area in controlling the photocatalytic reactivity of photocatalysts. Results also showed that the secondary particle size grew with time due mainly to particle aggregation. The photocatalytic rate constants decreased exponentially with increase in primary particle size. Primary particle size alone is able to predict the photocatalytic rate as it is closely related to the electro-optical properties of photocatalysts. # 2006 Published by Elsevier B.V. Keywords: Size effect; TiO 2 ; Photocatalyst; Particle size; 2-Chlorophenol; Size quantization effect; Electro-optical property 1. Introduction Since the discovery of photovoltaic property of titanium dioxide TiO 2 by Fujishima and Honda [1], great efforts have been focused on elucidating the electronic structure [2–5], catalytic reactivity [6–8] and surface property [9] of TiO 2 . Inexpensive and thermal-dynamically stable at room tempera- ture, this semiconductor material has been widely used in heterogeneous photocatalysis and proven to be capable of decomposing a host of organic pollutants such as phenolic compounds [10,11], metal ethylene diamine tetra acetate (EDTA) complexes [12,13], airborne microbes [14] and odorous chemicals [15]. Most of these studies involved ultraviolet (UV) photons as the major exciting light sources. Considering that there is only 5% of solar irradiation within the UV range, intuitively it is desirable to enhance the photo- catalytic performance of TiO 2 by enabling it to utilize photons from the near-visible to visible region. It has been suggested that this can be achieved by manipulating the particle size of photocatalyst [3,6,16] or doping the TiO 2 with foreign ions [17–19]. The initiation of a photocatalytic reaction requires a minimum photon energy that exceeds the band gap of the material in order to trigger the interband transition of electrons www.elsevier.com/locate/apcatb Applied Catalysis B: Environmental 68 (2006) 1–11 * Corresponding author. Tel.: +1 302 831 8428. E-mail address: [email protected] (C.P. Huang). 0926-3373/$ – see front matter # 2006 Published by Elsevier B.V. doi:10.1016/j.apcatb.2006.07.018

Upload: trinhxuyen

Post on 04-Jan-2017

222 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Size dependency of nanocrystalline TiO2 on its optical property and

www.elsevier.com/locate/apcatb

Applied Catalysis B: Environmental 68 (2006) 1–11

Size dependency of nanocrystalline TiO2 on its optical property and

photocatalytic reactivity exemplified by 2-chlorophenol

H. Lin a, C.P. Huang a,*, W. Li b, C. Ni b, S. Ismat Shah b,c, Yao-Hsuan Tseng d

a Department of Civil and Environmental Engineering, University of Delaware, Newark, DE 19716, United Statesb Department of Materials Sciences and Engineering, University of Delaware, Newark, DE 19716, United States

c Department of Physics and Astronomy, University of Delaware, Newark, DE 19716, United Statesd Energy and Environment Laboratories, Industrial Technology Research Institute, Hsinchu 301, Taiwan, ROC

Received 19 January 2006; received in revised form 7 July 2006; accepted 18 July 2006

Available online 1 September 2006

Abstract

Anatase TiO2 nanocrystallines (17–29 nm) were successfully synthesized by the metal–organic chemical vapor deposition method (MOCVD).

Moderate manipulation of system parameters of MOCVD can control the particle size. The electro-optical and photocatalytic properties of the

synthesized TiO2 nanoparticles were studied along with several commercially available ultra-fine TiO2 particles (e.g., 3.8–5.7 nm). The band gap

of the TiO2 crystallines was determined using the transformed diffuse reflectance technique according to the Kubelka–Munk theory. Results

showed that the band gap of TiO2 monotonically decreased from 3.239 to 3.173 eV when the particle size decreased from 29 to 17 nm and then

increased from 3.173 to 3.289 eV as the particle size decreased from 17 to 3.8 nm. The results of band gap change as a function of particle size

agreed well with what was predicted by the Brus’ equation, i.e., the effective mass model (EMM). However, results of the photocatalytic oxidation

of 2-chlorophenol (2-CP), showed that the smaller the particle size, the faster the degradation rate. This is attributed in part to the combined effect of

band gap change relative to the spectrum of the light source and the specific surface area (or particle size) of the photocatalysts. The change of band

gap due to particle size represents only a small optical absorption window with respect to the total spectrum of the light source, i.e., from 380 to

400 nm versus>280 nm. Consequently, the gain in optical property of the larger particles was severely compromised by their decrease in specific

surface area. Our results clearly indicated the importance of specific surface area in controlling the photocatalytic reactivity of photocatalysts.

Results also showed that the secondary particle size grew with time due mainly to particle aggregation. The photocatalytic rate constants decreased

exponentially with increase in primary particle size. Primary particle size alone is able to predict the photocatalytic rate as it is closely related to the

electro-optical properties of photocatalysts.

# 2006 Published by Elsevier B.V.

Keywords: Size effect; TiO2; Photocatalyst; Particle size; 2-Chlorophenol; Size quantization effect; Electro-optical property

1. Introduction

Since the discovery of photovoltaic property of titanium

dioxide TiO2 by Fujishima and Honda [1], great efforts have

been focused on elucidating the electronic structure [2–5],

catalytic reactivity [6–8] and surface property [9] of TiO2.

Inexpensive and thermal-dynamically stable at room tempera-

ture, this semiconductor material has been widely used in

heterogeneous photocatalysis and proven to be capable of

decomposing a host of organic pollutants such as phenolic

* Corresponding author. Tel.: +1 302 831 8428.

E-mail address: [email protected] (C.P. Huang).

0926-3373/$ – see front matter # 2006 Published by Elsevier B.V.

doi:10.1016/j.apcatb.2006.07.018

compounds [10,11], metal ethylene diamine tetra acetate

(EDTA) complexes [12,13], airborne microbes [14] and

odorous chemicals [15]. Most of these studies involved

ultraviolet (UV) photons as the major exciting light sources.

Considering that there is only 5% of solar irradiation within the

UV range, intuitively it is desirable to enhance the photo-

catalytic performance of TiO2 by enabling it to utilize photons

from the near-visible to visible region. It has been suggested

that this can be achieved by manipulating the particle size of

photocatalyst [3,6,16] or doping the TiO2 with foreign ions

[17–19].

The initiation of a photocatalytic reaction requires a

minimum photon energy that exceeds the band gap of the

material in order to trigger the interband transition of electrons

Page 2: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–112

between the lowest unoccupied molecule orbital (LUMO) and

the highest occupied molecule orbital (HOMO); that is, the

incident wavelength needs to be smaller than the wavelength of

the band gap threshold, lbg. Thus, it is speculated that reducing

the band gap of TiO2 can enhance its photocatalytic

performance through more efficient utilization of lower energy

photons. TiO2 has three distinct crystalline structures: rutile,

anatase, and brookite. Most studies on the photocatalytic

reactivity were conducted with either rutile or anatase, which

reported that the indirect band gap was 3.0 eV (or lbg 413 nm)

and 3.2 eV (or lbg 387 nm), respectively. Although rutile has a

lower band gap than anatase, it has been demonstrated that

anatase-structure TiO2 exhibits a better photocatalytic perfor-

mance than that of rutile [20–22]. This is attributed in part to a

wider optical absorption band and smaller electron effective

mass of rutile than those of anatase, which leads to higher

mobility of charge carriers in rutile than anatase [5]. The band

gap of TiO2 is commonly believed to be indirect. In contrast to

direct interband transition, indirect transition requires phonons

(lattice vibration) to compensate for the change in wave vector

during electron transition.

It has been reported that the band gap of semiconductor

crystalline is a function of the particle size [3,4,23]. Below a

certain threshold, the density of point/surface defects of

semiconductor crystalline increases with decrease in particle

size. Due to mild delocalization of molecular orbitals on the

surface, defects in the bulk semiconductor create deep and

shallow traps near the band edge of its electronic state, which

brings about reduction in band gap, that is, red-shift in

absorption spectrum [3,4]. When the size of semiconductor

particle decreases from its bulk to that of Bohr radius, e.g., the

first excitation state, the size quantization (Q-size) effect arises

due to the spatial confinement of charge carriers. Consequently,

electrons and holes in the quantum sized semiconductor are

confined in a potential well and do not experience the

delocalization that occurs in the bulk phase. Therefore, the band

gap of ultra-fine semiconductor particle increases with the

decrease in particle size when it is smaller than the band gap

minimum [3,4]. This phenomenon has been described by the

Brus’ effective-mass model (EMM) [3]. Size quantization

effect has been studied using various semiconductors including

CdS [24], HgSe, PbSe, CdSe [25], ZnO [26], Cd3S2 [27], and

TiO2 [6,16,28]. The reported Q-size effect of semiconductor

clusters appears to be between 1 and 12 nm. Results of these

studies were mostly obtained from liquid phase UV–vis

absorption spectrum, and few had focused on the photode-

gradation of organic compound.

A good photocatalyst must have high photon conversion

efficiency in addition to high specific surface area. In fact, the

primary particle size of photocatalysts determines both the

specific surface area and the photon conversion efficiency.

Although the size dependency of band gap has been studied

theoretically, only a few investigations have been conducted on

the change of band gap as a function of particle size using TiO2

[6,16,29]. Anpo et al. [6] studied the change of band gap of

Ti02over a wide range of particle sizes (e.g., 3.8–200 nm) and

found significant blue shifts of the absorption edge by 0.093 and

0.156 eV for rutile and anatase crystalline, respectively, when

particle size was less than 12 nm. Kormann et al. [16] observed

the quantum confinement effect upon illumination of TiO2

colloids (e.g., <3 nm) and reported a blue-shift by 0.15–

0.17 eV in absorption spectrum. However, Serpone et al. [29]

did not observe a quantum size effect in the particles size range

between 2.1 and 26.7 nm of TiO2. It must be mentioned that

these authors depended on UV–vis absorbance measurements

to estimate the band gap energy. Obviously, this method cannot

exclude the light scattering effect during absorbance measure-

ments, which would lead to an over-estimation of the

absorbance, especially for large aggregates. It is known that

the scattering efficiency is proportional to the fourth-power of

particle radius (e.g., r4) according to Rayleigh scattering theory

[30]. During UV–vis absorbance measurements, the size of

aggregates will increase with time in suspensions. Furthermore,

most of these UV–vis absorbance measurements were

conducted during the agglomeration/reflux stage of particle

preparation using the sol–gel process [6,16,29]. The presence

of precursor intermediates can contribute light absorbance also,

which, in turn, might affect the reliability and the interpretation

of results observed.

As mentioned above, the particle size can affect the

photocatalytic reactivity. A few studies have been conducted to

assess the relationship between particle size and photocatalytic

reactivity [6,22,30–38]. Anpo et al. [6] observed an increase in

quantum yield in the hydrogenation of CH3CCH when the

particle size of anatase TiO2 decreased from 11 to 3.8 nm and

concluded that it was caused by the quantum size effect. Maira

et al. [30] studied the photocatalytic degradation of trichlor-

oethylene (TCE) in gas phase upon TiO2 in the size range of 2.3

and 27 nm and reported an optimum particle size of 7 nm. They

further concluded that the lower reactivity at particle size less

than 7 nm was due to changes in structure and electronic

properties of the ultra-fine particles. Almquist and Biswas [31]

compared the effect of particle size using both flame

synthesized and commercial TiO2 particles, on the photode-

gradation of phenol. They reported an optimal particle size in

the range of 25–40 nm with respect to photo-degradation rate

and optical responses. However, the TiO2 samples consisted of

a mixture of anatase and rutile crystallines at various

proportions that results may be affected by the anatase/rutile

ratio in addition to particle size. Hao et al. [39] and Jang et al.

[33] studied the effect of anatase particle size on the

photodegradation of rhodamine B and methylenen blue dye.

They observed that the photocatalytic reactivity monotonically

increased with decrease in particle size in the range of 8–15 and

15–30 nm. However, the effect of size-dependent band gap

change and its corresponding excitation source spectrum was

not considered in their work. Nam et al. [34] studied the

photocatalytic degradation of TCE (tetrachloroethene) using

TiO2 (anatase) thin film and found an increase in the

photocatalytic degradation rate as the primary particle size

decreased from 37 to 25 nm. Zhang et al. [35] examined the

effect of anatase particle size on the oxidation of trichlor-

omethene, CHCl3, and reported that TiO2 particles at 11 nm

yielded the highest photonic efficiency on the oxidation of

Page 3: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–11 3

CHCl3. They reported that photocatalytic efficiency did not

increase monotonically with decrease in particle size; rather it

increased with an increase in surface combination of the

electron–hole pairs. Xu et al. [36] and Gerischer [37] have

experimentally and theoretically studied the effect of particle

size of on the photocatalytic efficiency of TiO2. Both

researchers concluded that the photocatalytic reactivity of

TiO2 increases with decrease in particle size. Unfortunately, the

TiO2 used in the studies conducted by Xu et al. [36] and

Gerischer [37] had a particle size mainly in the range of

micronmeters rather than nanometers. As indicated above,

there appears to be no agreement on the effect of particle size on

the photocatalytic activities of TiO2; the optimal particles size

reported has covered a rather wide range, e.g., between 3.8 and

40 nm. Another issue related to the effect of particle size on the

photocatalytic reaction is the primary versus the secondary

particle size. This is of particular importance when dealing with

aqueous systems, as particle aggregation is inevitable in the

water environment [38]. Will particle aggregation affect the

photocatalytic reactivity during the course of water treatment

process? Maira et al. [30] reported that the degradation of TCE

in the gas phase was affected by both the primary and the

secondary particle sizes. Little information with regard to the

importance of secondary particle size on photocatalytic

reactivity in aqueous solution is available. There is a need

for more systematic studies on the effect of particle size, both

primary and secondary, on photocatalytic reactions in aqueous

solutions.

The objectives of this study were (1) to elucidate the effect of

primary particle size on the electro-optical property of TiO2 in

terms of band gap changes, (2) to assess the effect of primary

and secondary particle size on the photocatalytic reactivity of

TiO2 exemplified by 2-chlorophenol (2-CP), and (3) to

determine whether any relationship exists between the optical

property and photocatalytic reactivity. In order to better

evaluate the effect of particle size on the optical property of

TiO2 with minimum optical interference (e.g., aggregation in

the aqueous phase), band gap measurements were made using

the transformed diffuse reflectance technique according to the

Kubelka–Munk theory [40–42]. It is expected that this

Table 1

TiO2 particle size (nm) analyses by various techniques

Particle TEMa BETb XRDc DLSd

A3.8 3.8 3.94 9.22 51.1

A4.9 4.9 5 10.71 25.9

A5.7 5.7 5.53 11.55 71.6

A12 12 10.56 15.1 –

A17 17 16.95 – 161.5

A20 20 20.06 17.07 92.8

A29 29 28.05 24.3 183.5

a Gaussian fitting calculated particle size based on TEM bright field images.b Calculated particle size based on BET results.c Scherrer equation calculated particle size based on XRD results [anatase(1 0 1d Dynlamic light scattering measured particle size right after the application ofe Dynamic light scattering measurements during the experiments (smaller clustef Dynamic light scattering measurements during experiments (larger cluster).g Average particle size calculated by the weighted PSD function.

technique will enable better correction of interferences in

absorbance measurements caused by particle light scattering

and other factors such as the presence of precursor

intermediates. Results of band gap measurements were then

compared with those predicted by the Brus’ EMM model. The

effect of particle size on the photocatalytic reactivity of

TiO2was assessed using a 2-chlorophenol (2-CP) probe over a

wide range of primary particle sizes, e.g., from 3.8 nm

(reported known quantization effect) to 29 nm (approximated

bulk size). The next step was to compare the effect of primary

and secondary particle size on the photocatalytic degradation of

hazardous organic compounds exemplified by 2-CP. Finally, the

band gap change was compared with the photocatalytic

degradation rate of 2-CP for assessing the overall effect of

particle size.

2. Materials and methods

2.1. TiO2 particle preparation

Pure anatase TiO2 particles with diameter ranging from 12 to

33 nm were prepared by metal–organic chemical vapor

deposition (MOCVD). Details of the system configuration

and operation have been reported previously [43]. Titanium

tetraisopropoxide—Ti[OCH(CH3)2]4 (TTIP), 97% purity, pur-

chased from Aldrich, was selected as the precursor. The liquid

TTIP was placed in a Pyrex thermal well. Argon (99.999%

purity) was used as the carrier gas. The precursor flow rate was

adjusted by controlling the temperature of the TTIP solution

and/or the flow rate of the Ar gas. Ar gas flow at 5–10 sccm was

purged through the thermal well to carry the precursor to the

reactor. TTIP was pre-heated to just below its boiling point of

220 8C to increase its volatility and deposition rate. The total

pressure of the reactor was controlled at 10–15 Torr by purging

O2 at 15–35 sccm of flow rate to the Ar/TTIP mixture in a

baffle. The mixture was directly introduced to the reaction

chamber for the formation of TiO2 particles. Stainless mesh was

selected as the substrate, which was triple cleaned by methanol

and de-ionized water prior to experiments. Particles were

collected from the stainless mesh substrate placed in the middle

DLSe DLSf DLSg Circularity

126.9 505.5 387.7 0.838 � 0.021

121.9 400.9 255.8 0.842 � 0.017

140.7 560.4 391.3 0.847 � 0.020

– – – 0.846 � 0.037

125.3 – 125.3 0.833 � 0.05

103.7 – 103.7 0.828 � 0.083

167.3 – 167.3 0.827 � 0.072

)].

ultrasonic disintegration at an energy intensity of 18 kJ/L.

r).

Page 4: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–114

of the reaction chamber and perpendicular to the direction of

the TTIP/Ar flow. The temperature of the furnace was kept at

600 8C to assure the formation of anatase structured TiO2

polycrystallines. The post-decarbonation time was 30 min,

which was performed by constant purging of O2 after the TTIP/

Ar valve was turned off. For particles smaller than 10 nm, two

samples (3.8 � 0.3 and 4.9 � 0.8 nm) were purchased from

Reade1 Inc., and one sample with a size of 5.7 � 0.3 nm was

provided by Hombikat Inc. In addition, P25, which contains

approximately 75% of anatase and 25% of rutile, provided by

Degussa Co. was also used for comparison. Table 1 lists the

important physical–chemical properties of the TiO2 samples

used in this study.

2.2. TiO2 surface characterization

Particles were characterized by various techniques, includ-

ing transmission electron microscopy (TEM), X-ray diffraction

(XRD), scanning electron microscopy (SEM), BET specific

surface area measurement, dynamic light scattering (DLS),

diffuse reflectance spectra, and X-ray photoelectron spectro-

scopy (XPS). The particle size and shape of MOCVD

synthesized TiO2 particles were determined based on TEM

(JEOL 2000 FX) results (insets in Fig. 1) and further

Fig. 1. Particle size distribution histograms based on TEM images. Inset—

TEM bright field image of TiO2 particles: (a) A29 and (b) A4.9.

reconfirmed by BET, SEM, and XRD measurements

(Table 1). For the sake of simplicity, we designate Ax as

anatase TiO2 particles with ‘‘x’’ nm in diameter. TiO2 particles

were immersed in acetone solution after ultrasonic treatment

for 5 min (Branson model 1510, 150 W, 40 kHz). The dispersed

TiO2 particles were then deposited onto the lacey-carbon-

coated copper grid. A JEOL 2000 equipped with a LaB 6 field

emission gun, system was operated at 2.5 � 10�5 Pa pressure.

An acceleration voltage of 200 kV was used. Filament current

was controlled between 111 and 113 mA.

SEM images were taken by a JEOL scanning electron

microscope, model JSM 7400F. The working distance was kept

between 2.8 and 3.1 mm, and the acceleration voltage was set at

3.0 keV. Example images of A17 and A23 synthesized TiO2

particles are shown in Fig. 2. EDX results (data not shown) did

not show any impurity in these TiO2 samples studied regardless

of their source of origin. Structural characterization was done

using XRD scan with Cu Ka radiation in a Rigaku D-max B

diffractometer equipped with a graphite crystal monochroma-

tor. The diffraction pattern obtained showed that only anatase

Fig. 2. Example SEM images of MOCVD synthesized TiO2 particles: (a) A17

and (b) A23.

Page 5: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–11 5

TiO2 crystalline formation (with neither rutile nor brookite

structure) was observed in any of the samples analyzed. The

crystalline size along the (h k l) profile was calculated based on

a high-resolution XRD scan following the Scherrer formula

(Eq. (1)), i.e.:

Dðh k lÞ ¼kl

b cos u(1)

where k is the shape factor, l the wavelength of X-ray of Cu

Ka radiation, b the full width at half maximum (FWHM) of the

(h k l) peak, and u is the diffraction angle. Results of XRD

survey scan are shown in Fig. 3a. It is obvious that only anatase

was present in the TiO2 samples studied. Based on the high-

resolution scan, the particle size was calculated using the

anatase(1 0 1) peak (Fig. 3b).

The specific surface was measured based on the Brunauer–

Emmett–Teller multiplayer nitrogen gas adsorption (BET)

theory (NOVA2000 Gas Sorption Analyzer, Quantachrome

Corp.). Samples were degassed and calcined at 300 K for 24 h

before adsorption experiments using N2 as the adsorbate. The

relative pressure (P/P0) was within the range of 0.05–0.35. The

average pore radius was also recorded.

Secondary particle size (aggregates) measurements were

conducted with a zetameter in dynamic light scattering mode

(DLS), Zetasizer, model 3000 HSA, Malvern Instruments Ltd.,

Fig. 3. XRD spectra of TiO2 samples: (a) survey spectrum and (b) high

resolution scan on anatase(1 0 1) peak region.

UK. Particles were prepared in solutions containing 50 mg/L of

2-CP at pH 10. The solution was ultrasonicated (18 kJ/L)

immediately prior to size measurements. Readings were taken

immediately after ultrasonication and 5 min after photocatalytic

experiments were started. The real refractive index and

imaginary refractive index for both the solvent and the analytes

were obtained according to the user’s manual for the instrument.

Diffuse reflectance spectra of the powders were measured on a

double beam UV–vis–NIR scanning spectrophotometer (Shi-

madzu UV-3101PC, Japan), which was equipped with a diffuse

reflectance accessory. A given amount of TiO2 powder was

uniformly pressed in the tablet (provided by Shimadzu) and

placed in the sample holder on integrated sphere for the

reflectance measurements. The reflectance data was converted to

the absorption coefficient F(R1) values according to the

Kubelka–Munk equation (Eq. (2)) [40–42], i.e.:

FðR1Þ ¼ð1� R1Þ2

2R1(2)

where F(R1) is equivalent to the absorption coefficient.

A SSI-M probe XPS system employing Al Ka exciting

radiation was used for the general survey and high resolution

scans. Ti 2p, O 1s, and C 1s photoelectron peaks were recorded

in the high-resolution mode. These peaks were used to

determine the relative composition of the nanoparticles and the

valence states of Ti. XPS results of MOCVD-prepared TiO2

were reported earlier [43]. In addition, a comparison of

particles from different sources in our study is shown in Fig. 4.

(Note: A carbon peak was present in Fig. 4. This was from the

background carbon tape and not that of the TiO2 sample.) Our

XPS results did not show any impurity for samples collected

from difference sources at a detection limit of 0.05 at.%. Based

on the above surface analyses, namely, XPS and EDX, all TiO2

particles studied, regardless of its origin, appeared to be

homogeneous and compatible with each other in terms of

crystal structure, morphology and purity. It suffices to say that

particle size was the only observable difference among these

TiO2 samples in presented study.

Fig. 4. XPS results of TiO2 particles from different sources: (1) A3.8 (Reade1

Inc.), (2) A5.7 (Hombikat Inc.), and (3) A29 (MOCVD prepared TiO2).

Page 6: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–116

Fig. 5. (a) Diffuse reflectance spectra of various TiO2 anatase particles accord-

ing to the Kubelka–Munk equation; (b) the transformed Kubelka–Munk func-

tion vs. energy of the excitation source.

2.3. Photocatalytic reactivity experiments

Photodegradation experiments were performed in a photo-

catalytic reactor system. This bench-scale system consisted of a

cylindrical Pyrex-glass jacketed cooling cell with a reflective

interior surface. The system was cooled and maintained at

298 K. Photon source was provided by a 100-W medium

pressure mercury lamp (Hanova, NJ, USA). TiO2 (at a

concentration of 10 mg/L) nanoparticles were ultrasonicated

(at 18 kJ/L) before the experiment was started. The target

organic compound was 2-chlorophenol (2-CP) (>99.9% purity

from Aldrich). Constant air purging was provided at a flow rate

of 100 sccm using a porous diffuser. The initial pH value of the

mixture solution was set at 10.0 to minimize the evaporation of

2-CP during experiments (note that the pKa of 2-CP is 8.52).

The residual concentrations of the parent and intermediate

compounds were measured using high-performance liquid

chromatography (HPLC) (Model HP 2100, Agilent, CA, USA).

The instrument was equipped with a Zorbax Eclipse XDB-C18

column (4.6 mm � 250 mm) from Agilent. The eluent con-

sisted of 60% acetonitrile and 40% phosphate (pH 3) buffer.

The system flow rate was maintained at 0.5 ml/min with total

injection volume of 10-ml. Spectra were acquired by a diode

array detector (l = 210 nm).

3. Results and discussion

3.1. Optical properties

From the diffuse reflectance plot (Fig. 5a), we can see that

A20, A17 and A23 have better light absorbance than A3.8,

A4.9, A5.7, and A29 in the range of wavelength (l) between

400 and 350 nm. The damping of the F(R1) for A3.8 and A4.9

was observed when the wavelength was below 360 nm. We

suspect that is due to the size quantization effect and defects

induced deep energy traps that are more discretely created

(compared to A5.7). The damping of the diffuse reflectance

may be caused by attenuation of excited electron de-excited

during the scanning of different wavelength. Thus, the

luminescence is created during the de-excitation and amplified

the signal of the reflectance. The absorption coefficient of an

indirect semiconductor near the absorption threshold can be

expressed as

a ¼ Biðhn� EgÞ2

hn(3)

where Eg is the band gap of indirect allowed transition (eV), h

the Planck’ s constant (J s), Bi the absorption constant, and n is

the frequency of the light (s�1). Therefore, a transformed

Kubelka–Munk function can be constructed by plotting

[F(R1)]0.5 against the energy of excitation source to obtain

the band gap of TiO2 particles (Fig. 5b) [41,44]. The band gaps

for A3.8, A4.9, A5.7, A17, A20, A23, and A29 were deter-

mined to be 3.289 � 0.02, 3.251 � 0.02, 3.275 � 0.02,

3.173 � 0.02, 3.179 � 0.02, 3.224 � 0.02, and 3.239 �0.02 eV, respectively. Results showed that the band gap

decreased (e.g., red shifted) to a certain minimum value

(i.e., critical size) with decrease in particle size from an

estimated bulk value of 29 nm. Further decrease in particle

size from the critical size (corresponding to band gap mini-

mum) caused the band gap to increase (i.e., blue-shift). A

plausible explanation for this change in band gap with respect

to the size of the particle is that the bulk defects induce

delocalization of molecular orbitals in the conduction band

edge (e.g., LUMO) and create shallow/deep traps in electronic

energy, in turn causing the red-shift of the absorption spectra.

When crystalline size decreased below its size at the band gap

minimum, the traps shifted to higher energy, which resulted in

blue shifting of the absorption spectra (e.g., size quantization

effect) [3,4]. Based on the Bras’ EMM model, the first excitonic

energy of semiconductor cluster, E*, can be expressed as a

function of particle size, as given by the following expression:

E� ffiEg þ�h2p2

2R2

�1

me

þ 1

mh

�� 1:8e2

eR(4)

where Eg is the band gap of the bulk semiconductor, the second

term is the energy induced by the quantum confinement effect,

Page 7: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–11 7

Fig. 6. Band gap shift vs. the particle size of TiO2. Solid circles are experi-

mental results. Lines are predicted by the Brus model at different reduced

effective mass of exitions, m (1/m = 1/me + 1/mh) and dielectric constant

k = 184.

Fig. 7. Photocatalytic degradation of 2-CP under UV radiation: (a) changes of

normalized concentration as a function of time; (b) least square best fitting for

the first-order reaction rate constant k (s�1). Symbols: P25 (solid squares), A29

(solid circles), A20: (open circles), A17 (open squares), A5.7 (solid diamonds),

A4.9 (open triangles), and A3.8 (open diamonds).

and the third term is the shift of energy due to columbic

attraction between electron and hole pairs. £ is the Plank’ s

constant (J s), R the radius of the cluster (m), me the effective

mass of the electron (kg), mh the effective mass of the hole (kg),

e the charge of electron (C), and e is the electric permittivity

(C2 N�1 m�2) of the material (a product of dielectric constant k

and permittivity in free space e0). As shown in Eq. (4), the band

gap increased as particle size decreased to below a certain

threshold. The shifts in band gap were significantly affected by

the reduced mass of exciton, i.e., m�1 ¼ m��1e þ m��1

h and the

dielectric constant of the material k. The reported effective

mass of the electron, m�e was in the range of 5–13me [16,29,45].

The effective mass of the hole, m�h was reported to be �2me

[16]; the dielectric constant of TiO2 anatase crystalline was

reported to be between 23 and 184 [29,46–49]. Eq. (4) was

plotted by using the reported values for the dielectric constant f

of 184 and effective mass of exitons (m�e) from 5 to 13me and m�hof 2me (Fig. 6). This was equivalent to m values of between 1.43

and 1.73me. We can see the decrease (left shift) in size of band

gap minimum and increase in band gap shift (DEg) when m

increases from 1.43 to 1.73. In contrast, decrease in dielectric

constant from 184 to 23 will result in greater reduction in band

gap minimum and significantly enlarge the band gap shift in the

negative direction (result not shown). By comparing the

observed band gap values to those calculated by the EMM

model, it is seen that the observed values agreed very well with

those predicted by the model.

3.2. Photoreactivity

If the energy of the incident photon is greater than the band

gap, photoexcitation will occur and yield the electron and hole

pairs. The trapped charge carriers are formed within a 20-ps

pulse but have a life time in the nanosecond (ns)-range at Ti3+

sites within the semiconductor bulk [50]. This mechanism has

been confirmed by direct and indirect in situ electron

paramagnetic resonance (EPR) measurements (e.g., examina-

tion of paramagnetic species on hydrated TiO2 surfaces)

[6,51,52]. Generated electrons are readily trapped on Ti4+ sites

and form Ti3+. The trapped electron can be readily scavenged

by oxygen [51]. Localized holes can be scavenged by either

reacting with hydroxide ions or through electron transfer with

water to form hydroxyl radicals, which are strong oxidants that

can oxidize organic substances nonselectively [10,31,51,52].

It is generally accepted that photogenerated hydroxyl

radicals are primarily responsible for the heterogeneous

degradation of organic compounds, such as 2-CP

[10,31,53,54]. Fig. 7 shows the photocatalytic reactivity of

2-CP over TiO2 particles at various particle sizes. Results

indicated that the degradation reaction followed pseudo first-

order kinetics regardless of the particle size (Table 2). It also

showed that the reaction rate constants increased with decrease

in primary particle size. This is in contrast to the band gap

measured (Fig. 5b), where a minimum band gap was observed

with a specific primary particle size of �17 nm. That is, the

particles with size in the range of 17–29 nm, exhibited larger

lbg values than the smaller ones.

Photoreactivity under UV irradiation showed that the

smaller the particles, the higher the 2-CP decomposition rate

Page 8: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–118

Table 2

List of the least square best fitting parameters for different primary particle sizes

Diametera Specific

surface areab

k � 1000c R2 QE � 1000d

P25 33 47.7 � 5 16.36 � 0.17 0.9565 12.75

A29 29 55.7 � 5 17.47 � 0.02 0.9997 13.49

A20 20 77.9 � 5 20.31 � 0.03 0.9995 13.70

A17 17 92.2 � 5 21.17 � 0.04 0.9993 13.71

A5.7 5.7 282.5 � 5 24.55 � 0.10 0.9974 13.87

A4.9 4.9 312.7 � 5 30.03 � 0.11 0.9982 14.02

A3.8 3.8 396.1 � 5 29.23 � 0.15 0.9901 13.95

a Diameter based on TEM, primary size (nm).b BET specific surface area (m2/g).c First-order reaction rate constant obtained by least squares best fitting (s�1).d Quantum efficiency at 180 min of experiment.

under otherwise identical experimental conditions. In our study,

commercially available P25 appeared to have the worst

degradation rate of 2-CP (k = 0.01636 s�1), whereas A3.8

yielded the highest rate (k = 0.02923 s�1). The specific surface

area of particles increases with decrease in particle size

(Table 1). Obviously, the advantage in photocatalytic activity

gained through the enhancement in optical response from the

smaller band gap exhibited by larger particles (e.g., 17–29 nm)

was severely compromised by the decrease in specific surface

area. Furthermore, the variation in photon absorption due to

size effect was lbg between 381 and 395 nm. Considering the

emission spectrum of the excitation source (Fig. 8), utilizable

photons are in the range between 280 nm (note that the Pyrex

glass cuts off the photon emission at 280 nm, as indicated in

Fig. 8) and lbg of TiO2. Clearly, the change of light absorption

band due to size effect (e.g., 381–395 nm) is relatively small

compared to the fully utilizable light spectrum (e.g., 280–

395 nm) available in this study. In addition, the quantum

efficiency of the photocatalytic reactivity appeared to have an

Fig. 8. Photonflux spectrum of the medium pressure mercury lamp used. The

two vertical lines, between 381 and 395 nm, indicate the band gap changes

observed of the TiO2 samples used in this study. Generally, the band gap of the

bulk TiO2 was around 387.2 nm. The band gap red-shifted to ca. 395 nm as the

particle size decreased to about 17 nm then the band gap blue-shifted to ca.

381 nm as the particle size continued to decrease to about ca. 3.8 nm. Inset is the

plot of particle size vs. its corresponding wavelength of band gap threshold lbg

(nm).

exponential relationship to the incident wavelength near the

optical adsorption edge. Zang et al. [55] reported that the

incident photon conversion efficiency (IPCE) could vary by an

order of magnitude comparing excitation wavelength at 400

and 550 nm during the photocatalytic degradation of 4-

chlorophenol (4-CP). Puma and Yue [54] observed that the

reaction rate of 2-CP increased by 3.9 times (1.73 mM min�1

versus 6.78 mM min�1) when the system was under UV-A

versus UV-ABC (i.e., the full UV spectrum) irradiation. These

observations imply that even when the incident photon energy

is higher than the band gap of the photocatalyst, quantum

efficiency varies greatly below the wavelength of the band gap

threshold, lbg. This variation is especially large when the

incident wavelength is near the adsorption edge. Thus, size-

effect-enhanced optical absorption (381–395 nm in our study)

is likely to contribute an unnoticeable improvement on the

photocatalytic degradation of the 2-CP. As a result, it is not

surprising to see photocatalysts with lower band gap and

smaller specific surface area yield lower photodegradation

efficiency than those with higher band gap and larger specific

surface area. The quantum yield (F) was computed based on

Eq. (5) [48]:

Fl ¼mole of 2-CP degraded

Einstein of incident photons(5)

Fig. 9 shows the apparent quantum yield versus the primary

particle size at 180 min. It is seen that the quantum yield

increases almost monotonically with decreases in primary

particle size. Again, this is due to the advantage of the large

specific surface area of the ultra-fine particles, the spectrum of

the excitation source, and the wavelength-dependent nature of

photocatalytic reactivity.

In photocatalytic slurry systems, two major factors will

greatly impact the reaction: (a) specific surface area and (b) band

gap. The specific surface area determines the available sites for

reactions to take place, whereas, the band gap of the

semiconductor will define the amount of photons that are

available for quantum conversion. Under our experimental

conditions, results showed that ultra-fine particles (e.g.,<10 nm)

Fig. 9. Quantum yield as a function of primary particle size.

Page 9: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–11 9

Fig. 10. Secondary particle size measurements using the DLS method: (a)

measured immediately after the ultrasonic treatment; (b) measured 5 min after

the experiment was started.

have specific surface area nearly one order of magnitude larger

than that of larger particles (e.g., >10 nm). However, the ultra-

fine particles only exhibited a reaction rate roughly by a factor of

two and slightly higher apparent quantum efficiency (F) in

180 min than that of the larger particles. It seems that several

factors might offset the surface area advantage of ultra-fine

particles: (1) size quantization effect (e.g., less than 10 nm in our

case), (2) increase in surface electron–hole recombination, (3)

particle aggregation, and (4) spectrum of the excitation source.

Quantization effect yields a larger band gap as the particle size

decreases. As a result a light source with higher energy, e.g., blue-

shifted, is required to separate the excitons [3,4]. When the

particle size decreases, the density of recombination center

increases which encourages holes and electrons recombination

[35,38]. As discussed in Section 3.3, ultra-fine particles (e.g.,

<10 nm) can undergo rapid flocculation, which decreases the

availability of active surface sites. Finally, the energy band of

excitation source is also important to take advantage of the band

gap reduction due to size effect.

3.3. Effect of secondary particle size

Regardless of the primary particle size, it is inevitable that

there will be particle aggregation in aqueous solutions due to

factors such as the charge density/potential of the particle

surfaces and van der Waals forces. The effect of secondary

particle size (aggregates) was also assessed in this study. The

secondary particle sizes (aggregation) of TiO2 particles were

determined using dynamic light scattering (DLS) measure-

ments (Fig. 10) under two conditions: (1) right after the

ultrasonic treatment at an energy intensity of 18 kJ/L (herein

designated as ‘‘immediate’’ secondary particle size) and (2)

5 min (herein designated as ‘‘5-min’’ secondary particle size)

after the experiment was started. Fig. 10a shows the particle

size distribution (PSD) of the ‘‘immediate’’ secondary particles.

Results indicated the PSD was mono-dispersed. The average

particle size of the ‘‘immediate’’ secondary particles increased

from 3.8–29 to 51.1–183.5 nm, respectively, an increase by 13

to 6 times. However, as indicated in Fig. 10b, for the ‘‘5-min’’

secondary particles, the PSD was mono-dispersed when the

primary particle size was >10 nm and binomially distributed

when the primary particle size was <10 nm. If one calculates

the average ‘‘5-min’’ secondary particle size using the weighted

PSD function, that is, summing up the product of the particle

size of each size class and its corresponding distribution

frequency, the ‘‘5-min’’ secondary particle size increased from

3.8–29 to 387.7–167.3 nm, respectively, a 100 to 5 times

increase in particle size. This is expected as the flocculation of

smaller particles generally proceeds rapidly according to the

perikinetic flocculation rate expression [56]. During perikinetic

flocculation, Brownian diffusion is the major driving force. The

perikinetic flocculation rate is a second order expression with

respect to particle number concentration. Therefore, at identical

mass concentration, e.g., 10 mg/L, the smaller the primary

particles, the faster they aggregated. As the particle size

increased, perikinetic flocculation was overtaken by orthoki-

netic flocculation, which is driven by the extent of mixing and is

a first-order rate equation with respect to particle number

concentration. Our results agreed well with what would be

expected from the general principle of flocculation kinetics.

Fig. 11 shows the pseudo first-order 2-CP degradation rate

constants as a function of the primary and the secondary

particle size. It is seen that the photoreactivity decreased

exponentially with increase in primary particle size. As

indicated above, the primary particles underwent immediate

aggregation upon dispensed into the 2-CP solution at an ionic

strength of ca. 10�3 to 10�2 M. Generally the pseudo first rate

constant also decreased exponentially with the ’’immediate‘‘

secondary particle size. The rate constants remained relatively

unchanged at high values as the ‘‘5-min’’ secondary particle

size increased; the rate then decreased sharply as the ‘‘5-min’’

secondary particle size approached to a value of ca. 400 nm.

Further increase in the ‘‘5-min’’ secondary particle size resulted

in continuing decrease in rate constants. Results imply that

photocatalytic reactivity of each size-class of TiO2 particles

was better predicted by the primary particle size and the

‘‘immediate’’ secondary particle size. The specific surface area

advantage in terms of promoting photocatalytic reactions was

preserved to some extent during the aggregation process,

especially for the ultra-fine particles that underwent fast

flocculation. Our results agreed well with Maira et al. [30] who

Page 10: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–1110

Fig. 11. Photocatalytic reaction rate constants as a function of particle size.

TEM measured primary particle size (solid squares); size measured immedi-

ately after the ultrasonic treatment (open circles); size measured at 5 min after

the experiment was started (solid circles). Ultrasound energy = 18 kJ/L. Note

that the flocculation reaction occurred rapidly when the primarily particle size

was less than 10 nm. Immediately upon dispersion of the nano-TiO2 particles,

particle size grew by 6–13 times. The particle size continued to grow by 5–100

times for the ultra-fine particles (e.g., <10 nm) as the reaction proceeded.

reported similar effect of secondary particle size, immediately

measured, on the photocatalytic reaction rate in gas phase. No

information has been reported on the effect of secondary

particle size (e.g., ‘‘5-min’’ or longer) on photocatalytic

activities in aqueous solutions. Under otherwise similar

conditions, smaller particles (e.g., <10 nm) have the specific

surface area advantages but an increase in band gap which

limits their accessibility to the excitation energy due to blue-

shift in absorption spectrum. In aqueous solutions, although the

primary particle size and ‘‘immediate’’ secondary particle size

are able to predict the effect of particle size on the

photocatalytic reactivity in aqueous solutions, the ‘‘5-min’’

or longer time secondary particle size better represented the

actual temporal state of particles. Future research is needed to

couple the kinetics of particle flocculation and the photo-

catalytic reaction as to elucidate the total effect of particle size

on photocatalytic reactivity in aqueous solutions.

4. Conclusion

TiO2 crystallines of different particle size were successfully

synthesized by the MOCVD method (particle size ranged

between 12 and 29 nm). The particle size of TiO2 can be simply

manipulated by moderate control of the O2/precursor flow rate

and system pressure. Results from the transformed Kubelka–

Munk equation showed that the band gap of the TiO2

nanoparticles was a function of the primary particle size.

When the TiO2 particle size decreased from its bulk (i.e.,

=29 nm) to 17 nm, the band gap decreased. In contrast, for

ultra-fine TiO2 particles (e.g., 3.8, 4.9, and 5.7 nm), the band

gap increased by 0.05, 0.012 and 0.036 eV, respectively. This

could be attributed to both the delocalization of molecular

orbitals, which in turn creates energy traps and surface states on

the band edge, and the size quantization effect reported by other

researchers [3,6,16,30]. Obtained band gap values fitted the

Brus’ s EMM model well within the particle range of less than

30 nm. Although TiO2 particles in the size range of 17–23 nm

have lower band gap energy (0.01–0.06 eV lower than that of

the bulk), results showed that the photocatalytic reactivity of

TiO2 colloid in aqueous suspension increased with the decrease

in the primary particle size. This can be attributed in part to the

large specific surface area of smaller particles. Additionally,

UV lamp spectrum in the range of UV-B (290 nm > l >320 nm) and full UV-A (400 nm > l > 320 nm) provides

sufficient energy for exciton separation of higher band gap

energy in particles (e.g., A3.8, A4.9, and A5.7). We therefore

suggest that unless a well-confined light source is provided

(e.g., l = 380–400 nm), TiO2 particles in the size range of 17–

29 nm cannot perform better in terms of photocatalytic

reactivity than those in the 3.8–5.7 nm size range, under

otherwise identical experimental conditions. Furthermore, even

in the presence of light at wavelength between 380 and 400 nm,

the photocatalytic performance of large particles (e.g., 17–

29 nm) will be compromised due to their relatively low specific

surface area. Our results clearly indicated that particle

aggregation did take place readily in aqueous solutions. Unless

a constant energy input is provided throughout the entire

experiment, there is no way to maintain the primary particle

size. In light of the rapid aggregation among the ultra-fine

particles, especially those that are smaller than 10 nm, primary

particle size alone was able to predict the photocatalytic

reaction. Intuitively, one would reason that the primary particle

size holds key to the electro-optical property and photocatalytic

reactivity of photocatalysts due mainly to the change in band

gap energy (e.g., result in shifts in absorption spectrum) and its

specific surface area.

Acknowledgement

The authors wish to acknowledge our two anonymous

reviewers for their excellent comments. This work was

supported by a NSF grant NIRT #0210284.

References

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

[2] R. Asahi, Y. Taga, W. Mannstadt, A.J. Freeman, Phys. Rev. B 61 (2000)

7459–7465.

[3] L. Bras, J. Chem. Phys. 80 (1984) 4403–4409.

[4] L. Bras, J. Phys. Chem. 90 (1986) 2555–2560.

[5] S.D. Mo, W.Y. Ching, Phys. Rev. B 51 (1995) 13023–13032.

[6] M. Anpo, T. Shima, S. Kodama, Y. Kubokawa, J. Phys. Chem. 91 (1987)

4305–4310.

[7] A. Hagfeldt, H. Lindstrom, S. Sodergren, S.E. Lindquist, J. Electroanal.

Chem. 381 (1995) 39–46.

[8] B. Oregan, M. Gratzel, Nature 353 (1991) 737–740.

[9] K. Bourikas, T. Hiemstra, W.H. Van Riemsdijk, Langmuir 17 (2001) 749–

756.

[10] I. Ilisz, A. Dombi, K. Mogyorosi, A. Farkas, I. Dekany, Appl. Catal. B:

Environ. 39 (2002) 247–256.

[11] S. Nevim, H. Arzu, K. Gulin, C. Zekiye, J. Photochem. Photobiol. A:

Chem. 146 (2002) 189–197.

[12] S.V. Muhammad, A.P. Davis, Water Res. 34 (2000) 952–964.

[13] J.K. Yang, A.P. Davis, Environ. Sci. Technol. 35 (2001) 3566–3570.

Page 11: Size dependency of nanocrystalline TiO2 on its optical property and

H. Lin et al. / Applied Catalysis B: Environmental 68 (2006) 1–11 11

[14] S. Takashi, K. Yoshiyuki, T. Masahito, Biochem. Eng. J. 14 (2003) 149–152.

[15] M.C. Canela, R.M. Alberici, W.F. Jardim, J. Photochem. Photobiol. A:

Chem. 112 (1998) 73–80.

[16] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, J. Phys. Chem. 92 (1988)

5196–5201.

[17] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001)

269–271.

[18] W. Choi, A. Terrain, M.R. Hoffmann, J. Phys. Chem. 98 (1994) 13669–

13679.

[19] T. Umebayashi, T. Yamaki, H. Itoh, J. Phys. Chem. Solids 63 (2002) 1909–

1920.

[20] J. Augustynski, Electrochim. Acta 38 (1993) 43–46.

[21] X.H. Tang, Y. Zhang, G.F. Yin, D.L. Zhou, H. Liu, C.Q. Zheng, Rare Met.

Mater. Eng. 33 (2004) 864–868.

[22] K. Tanaka, M.F.V. Lapule, T. Hisanaga, Chem. Phys. Lett. 187 (1991) 73–76.

[23] J.P. Rino, N. Studart, Phys. Rev. B 59 (1999) 6643–6649.

[24] M. Haase, H. Weller, A. Henglein, J. Phys. Chem. 92 (1988) 4706–4712.

[25] J.M. Nedeljkovic, M.T. Nenadovic, O.I. Micic, A.J. Nozik, J. Phys. Chem.

90 (1986) 12–13.

[26] U. Koch, A. Fojtik, H. Weller, A. Henglein, Chem. Phys. Lett. 122 (1985)

507–510.

[27] A. Fojtik, H. Weller, A. Henglein, Chem. Phys. Lett. 120 (1985) 552–554.

[28] L. Kavan, T. Stoto, M. Gratzel, D. Fitzmaurice, V. Shklover, J. Phys.

Chem. 97 (37) (1993) 9493–9498.

[29] N. Serpone, D. Lawless, R. Khairatdinov, J. Phys. Chem. 99 (1995)

16646–16654.

[30] A.J. Maira, K.L. Yeung, C.Y. Lee, P.L. Yue, C.K. Chan, J. Catal. 192

(2000) 185–196.

[31] C.B. Almquist, P. Biswas, J. Catal. 212 (2002) 145–156.

[32] J.A. Byrne, B.R. Eggins, P.S.M. Dunlop, S. Linquette-Mailley, Analyst

123 (1998) 2007–2012.

[33] H.D. Jang, S.J. Kim, S.K. Kim, J. Nanoparticle Res. 3 (2001) 141–147.

[34] H.J. Nam, T. Amemniya, M. Murabayashi, K. Itoh, J. Phys. Chem. B 108

(2004) 8254–8259.

[35] Z. Zhang, C.C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102 (1998)

10871–10878.

[36] N. Xu, Z. Shi, Y. Fan, J. Dong, J. Shi, M.Z.C. Hu, Ind. Eng. Chem. Res. 38

(1999) 373–379.

[37] H. Gerischer, Electrochim. Acta 40 (1995) 1277–1281.

[38] M.A. Grela, A.J. Colussi, J. Phys. Chem. 100 (1996) 18214–18221.

[39] W.C. Hao, S.K. Zheng, C. Wang, T.M. Wang, J. Mater. Sci. Lett. 21 (2002)

1627–1629.

[40] G.F.A. Kortum, Reflectance Spectroscopy: Principles, Methods, Applica-

tions, New York, 1969.

[41] G. Burgeth, H. Kisch, Coord. Chem. Rev. 230 (2002) 41–47.

[42] E.L. Simmons, Appl. Opt. 14 (1975) 1380–1386.

[43] W. Li, M. Sung, C.P. Huang, I. Shah, J. Vac. Sci. Technol. B 20 (2002)

2303–2308.

[44] S. Sakthivel, H. Kisch, Angew. Chem. Int. Ed. 42 (2003) 4908–4911.

[45] J. Pascual, J. Camassel, H. Mathieu, Phys. Rev. B 18 (1978) 5606–5614.

[46] W. Dong, S.P. Ruan, X.D. Zhang, W.B. Guo, C.X. Liu, S. Zhang, C.P. Jia,

J.X. Pan, W.Y. Chen, Chem. Res. Chin. Univ. 19 (2003) 347–349.

[47] M. Es-Souni, I. Oja, M. Krunks, J. Mater. Sci. Mater. Electron. 15 (2004)

341–344.

[48] I. Oja, A. Mere, M. Krunks, C.H. Solterbeck, M. Es-Souni, Solid State

Phenom. 99/100 (2004) 259–262.

[49] R. VandeKrol, A. Goossens, J. Schoonman, J. Electrochem. Soc. 144

(1997) 1723–1727.

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

[51] R.F. Howe, M. Gratzel, J. Phys. Chem. 91 (1987) 3909–3960.

[52] C.D. Jaeger, A.J. Brard, J. Phys. Chem. 83 (1979) 3146–3152.

[53] C.D. Dong, C.P. Huang, Advances in Chemistry Series, vol. 244, Am.

Chem. Soc., Washington, DC, 1995, pp. 291–313 (Chapter 15).

[54] G.L.L. Puma, P.L. Yue, Ind. Eng. Chem. Res. 41 (2002) 5594–5600.

[55] L. Zang, W. Macyk, C. Lange, W.F. Maier, C. Antonius, D. Meissner, H.

Kisch, Chem. Eur. J. 6 (2000) 379–384.

[56] P.C. Hiemenz, R. Rajagopalan, Principles of Colloid and Surface

Chemistry, Marcel Dekker, Inc., 1997.