8
Synthesis of nanocrystalline photocatalytic TiO 2 thin films and particles using sol–gel method modified with nonionic surfactants Hyeok Choi a , Elias Stathatos b , Dionysios D. Dionysiou a, * a Department of Civil and Environmental Engineering, University of Cincinnati, Cincinnati, OH 45221-0071, USA b Engineering Science Department, University of Patras, GR-26500 Patras, Greece Received 7 June 2005; received in revised form 6 December 2005; accepted 14 December 2005 Available online 7 February 2006 Abstract A simple sol– gel route has been developed for the preparation of nanocrystalline photocatalytic TiO 2 thin films and particles at 500 -C. The synthesis involved a novel chemistry method employing nonionic surfactant molecules as a pore directing agent along with acetic acid-based sol – gel route without direct addition of water molecules. This study investigated the effect of surfactant type and concentration on the homogeneity, morphology, light absorption, dye adsorption and degradation, and hydrophilicity of TiO 2 films as well as on the structural properties of the corresponding TiO 2 particles. The method resulted in the synthesis of mesoporous TiO 2 material with enhanced structural and catalytic properties including high surface area, large pore volume, pore size controllability, small crystallite size, enhanced crystallinity, and active anatase crystal phase. The prepared TiO 2 thin films were super-hydrophilic and possessed thermally stable spherical bicontinuous mesopore structure with highly interconnected network. Highly porous TiO 2 films prepared with polyethylene glycol sorbitan monooleate surfactant exhibited four times higher photocatalytic activity for the decoloration of methylene blue dye than the nonporous control TiO 2 films prepared without the surfactant. This sol– gel method modified with surfactant templates is useful in the preparation of nanostructured anatase TiO 2 thin films with high photocatalytic activity and desired pore structure. D 2005 Elsevier B.V. All rights reserved. Keywords: Titanium oxide; Photocatalysis; Nanostructures; Nanomaterials; Sol – gel, Thin films; Particles; Surfactant; Dip-coating 1. Introduction The synthesis of nanocrystalline anatase TiO 2 material with high surface area has accelerated its widespread use in environmental remediation such as photocatalytic destruction of toxic organic compounds and inactivation of microorgan- isms in water and air [1–7]. In addition to the intrinsic photocatalytic activity of TiO 2 , which is directly related with its crystal properties, the structural properties of porous TiO 2 catalyst such as its surface area, porosity, and pore size and distribution are also of importance because of their potential role in enhancing the light absorbance of TiO 2 catalyst and the accessibility of reactants to the active catalytic sites. An interesting method to fabricate highly porous materials with desired pore structure and size for target-specific applications is the use of amphiphilic organic molecules such as surfactants and block copolymers as pore directing agents in sol–gel methods [8–12]. Ionic surfactants such as alkyl phosphate, dodecylamine, and cetyltrimethylammonium chloride were initially used in the synthesis due to the strong and well- organized incorporation of titania inorganic framework onto surfactant micelles by electrostatic interactions [9,11,12]. However, the use of ionic surfactants showed limited potential for such applications since the strong electrostatic binding force makes it difficult to remove the templates completely through extraction methods and even heat treatment at high temperature. Common sol–gel methods employing direct addition of water molecules in a sol can lead to the immediate precipitation of amorphous particles with uncontrolled structure due to the rapid hydrolysis and condensation reactions between the highly reactive alkoxide titanium precursors and water. During the past decade, an interesting synthetic route of the Ti–O–Ti network in the absence of externally added water molecules has been studied using acetic acid as a titania sol modifier in 0040-6090/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.tsf.2005.12.217 * Corresponding author. Tel.: +1 513 556 0724; fax: +1 513 556 2599. E-mail address: [email protected] (D.D. Dionysiou). Thin Solid Films 510 (2006) 107 – 114 www.elsevier.com/locate/tsf

Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

Embed Size (px)

Citation preview

Page 1: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

w.elsevier.com/locate/tsf

Thin Solid Films 510 (

Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles

using sol–gel method modified with nonionic surfactants

Hyeok Choi a, Elias Stathatos b, Dionysios D. Dionysiou a,*

a Department of Civil and Environmental Engineering, University of Cincinnati, Cincinnati, OH 45221-0071, USAb Engineering Science Department, University of Patras, GR-26500 Patras, Greece

Received 7 June 2005; received in revised form 6 December 2005; accepted 14 December 2005

Available online 7 February 2006

Abstract

A simple sol–gel route has been developed for the preparation of nanocrystalline photocatalytic TiO2 thin films and particles at 500 -C.The synthesis involved a novel chemistry method employing nonionic surfactant molecules as a pore directing agent along with acetic acid-based

sol–gel route without direct addition of water molecules. This study investigated the effect of surfactant type and concentration on the

homogeneity, morphology, light absorption, dye adsorption and degradation, and hydrophilicity of TiO2 films as well as on the structural

properties of the corresponding TiO2 particles. The method resulted in the synthesis of mesoporous TiO2 material with enhanced structural and

catalytic properties including high surface area, large pore volume, pore size controllability, small crystallite size, enhanced crystallinity, and active

anatase crystal phase. The prepared TiO2 thin films were super-hydrophilic and possessed thermally stable spherical bicontinuous mesopore

structure with highly interconnected network. Highly porous TiO2 films prepared with polyethylene glycol sorbitan monooleate surfactant

exhibited four times higher photocatalytic activity for the decoloration of methylene blue dye than the nonporous control TiO2 films prepared

without the surfactant. This sol–gel method modified with surfactant templates is useful in the preparation of nanostructured anatase TiO2 thin

films with high photocatalytic activity and desired pore structure.

D 2005 Elsevier B.V. All rights reserved.

Keywords: Titanium oxide; Photocatalysis; Nanostructures; Nanomaterials; Sol–gel, Thin films; Particles; Surfactant; Dip-coating

1. Introduction

The synthesis of nanocrystalline anatase TiO2 material with

high surface area has accelerated its widespread use in

environmental remediation such as photocatalytic destruction

of toxic organic compounds and inactivation of microorgan-

isms in water and air [1–7]. In addition to the intrinsic

photocatalytic activity of TiO2, which is directly related with its

crystal properties, the structural properties of porous TiO2

catalyst such as its surface area, porosity, and pore size and

distribution are also of importance because of their potential

role in enhancing the light absorbance of TiO2 catalyst and the

accessibility of reactants to the active catalytic sites. An

interesting method to fabricate highly porous materials with

desired pore structure and size for target-specific applications is

the use of amphiphilic organic molecules such as surfactants

0040-6090/$ - see front matter D 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.tsf.2005.12.217

* Corresponding author. Tel.: +1 513 556 0724; fax: +1 513 556 2599.

E-mail address: [email protected] (D.D. Dionysiou).

and block copolymers as pore directing agents in sol–gel

methods [8–12]. Ionic surfactants such as alkyl phosphate,

dodecylamine, and cetyltrimethylammonium chloride were

initially used in the synthesis due to the strong and well-

organized incorporation of titania inorganic framework onto

surfactant micelles by electrostatic interactions [9,11,12].

However, the use of ionic surfactants showed limited potential

for such applications since the strong electrostatic binding

force makes it difficult to remove the templates completely

through extraction methods and even heat treatment at high

temperature.

Common sol–gel methods employing direct addition of

water molecules in a sol can lead to the immediate precipitation

of amorphous particles with uncontrolled structure due to the

rapid hydrolysis and condensation reactions between the highly

reactive alkoxide titanium precursors and water. During the

past decade, an interesting synthetic route of the Ti–O–Ti

network in the absence of externally added water molecules has

been studied using acetic acid as a titania sol modifier in

2006) 107 – 114

ww

Page 2: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

H. Choi et al. / Thin Solid Films 510 (2006) 107–114108

alcohol solvent [13–17]. In such a synthesis method, the

following basic steps are involved: (1) replacement of alkoxy

groups in the titania organic precursor with acetate groups,

resulting in the formation of alcohol, (2) esterification reaction

between alcohol and acetic acid to form water molecules, and

(3) slow hydrolysis reaction caused by water released from the

esterification reaction or direct condensation reaction of

acetate-bonded titanium [13,14].

Thus, the presence of surfactant molecules in the sol is

expected to play a crucial role in synthesizing tailor-designed

TiO2 catalytic materials and further reducing the hydrolysis

and condensation reaction rates due to the capping effect of

surfactants around the titania precursor [17,18]. This might

induce a highly porous titania inorganic network. However,

many research studies have focused on either the surfactant

self-assembling techniques or the acetic acid assisted sol–gel

strategies. Little work has been devoted to understand the role

of surfactants in the acetic acid assisted sol–gel method on the

formation of nanocrystalline TiO2 material and investigate its

structural and photocatalytic properties. In addition to the

porous structure and catalytic activity of TiO2 films, sol

stability and film homogeneity are also important factors in

such fabrication procedures.

In order to overcome these concerns and challenges, this

research deals with the preparation of highly efficient

nanostructured photocatalytic TiO2 thin films and particles

via an acetic acid-assisted sol–gel method employing nonionic

surfactants. Systematic study is conducted to elucidate the

effect of surfactant type and concentration on the structural and

catalytic properties of TiO2 material as well as on the physical

properties of TiO2 thin films including homogeneity, repro-

ducibility, thickness, and hydrophilicity.

2. Experimental details

2.1. Sol synthesis

The organic molecules used as pore directing agents were

representative nonionic long chain surfactants including Tween

20 (T20, polyethylene glycol sorbitan monolaurate), Tween 80

(T80, polyethylene glycol sorbitan monooleate), and Triton X-

100 (X100, polyethylene glycol tert-octylphenyl ether) pur-

chased from Aldrich. Compared to other commonly used toxic

and ionic templating agents, these organics are relatively

inexpensive, biodegradable, non-toxic, and easily removable.

Such large amphiphilic molecules exhibit the existence of

ordered mesophase and the ability to synthesize tailor-designed

porous TiO2 catalytic materials [19,20]. Each surfactant was

dissolved in isopropanol (i-PrOH, Fisher). Before adding

alkoxide precursor, acetic acid (Fisher) was added into the

solution for the esterification reaction with alcohol. Then,

titanium tetraisopropoxide (TTIP, Aldrich) was added under

vigorous stirring. The molar ratio of surfactant/i-PrOH/acetic

acid/TTIP was R:45:6:1, where the surfactant concentration R

was varied in the range from 0.0 to 3.0. Regardless of

surfactant addition, the sol was transparent, homogeneous,

and stable.

2.2. Formation of TiO2 thin films and particles

For the synthesis of immobilized TiO2 thin films,

borosilicate glass (Micro slide, Gold Seal) substrate with an

effective surface area of 10 cm2 was cleaned with water

followed by acetone. The substrate was dip-coated with the

sol using a homemade dip-coating apparatus at a withdrawal

rate of 12.8 cm/min. After coating, the films were dried at

room temperature for 1 h, calcined at a ramp rate of 3 -C/min

in a programmable furnace (Paragon HT-22-D, Thermcraft),

stayed at 500 -C for 15 min, and cooled down naturally.

This procedure was repeated three times. For convenience, the

abbreviations, ‘‘filmsurfactant’’ and ‘‘filmcontrol’’ denote TiO2

film prepared with surfactant and control TiO2 film prepared

without surfactant, respectively. Because of the difficulty in

direct characterization of the porosity and crystal structure of

small quantity of TiO2 films immobilized on glass support,

material characterization was carried out on the corresponding

TiO2 particles obtained from thick films. Even though the

properties of TiO2 particles are not exactly the same as those

of TiO2 films, this technique is useful for quickly examining

and comparing the effect of sol conditions on the structural

properties of the final material [21,22]. For TiO2 powder

characterization, the sol was spread on the glass substrate,

dried, and heat-treated at 500 -C for 1 h (instead of 15 min as

in making thin films) to thoroughly remove all the organics,

resulting in the formation of a thick film. The TiO2 particles

were collected by scraping the thick films and ground for

further characterization.

2.3. Materials characterization

A Kristalloflex D500 diffractometer (Siemens) with Cu Ka

(k =1.5406 A) radiation was employed for X-ray diffraction

(XRD) study of the TiO2 catalyst. A porosimetry analyzer

(Tristar 3000, Micromeritics) was used to investigate structural

characteristics of TiO2 material including Brunauer, Emmett,

and Teller (BET) surface area, porosity, and pore size and

distribution after purging samples with nitrogen gas for 2 h at

150 -C using Flow prep 060 (Micromeritics). A UV–Vis

spectrophotometer (Hewlett Packard 8452A) was used to

measure the UV–Visible light absorption of TiO2 films for

determining their band gap energy and evaluating their light

utilization. For examining the morphology of TiO2 nanostruc-

ture, a JEM-2010F (JEOL) high-resolution transmission

electron microscope (HR-TEM) with field emission gun at

200 kV was used. The samples were dispersed in methanol

(HPLC grade, Pharmco) using an ultrasonicator (2510R-DH,

Bransonic) for 5 min and fixed on a carbon-coated copper grid

(LC200-Cu, EMS). An environmental scanning electron

microscope (ESEM, Philips XL 30 ESEM-FEG) at accelerat-

ing voltage of 10 K was used to measure the thickness of films

and examine film homogeneity. Elemental composition

analysis of TiO2 materials was performed using energy

dispersive X-ray spectroscope (EDX, Oxford Isis) connected

to the HR-TEM and ESEM. For measuring the contact angle

of the film surface, the sessile drop method was employed

Page 3: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

H. Choi et al. / Thin Solid Films 510 (2006) 107–114 109

using a static angle goniometer (Rame-Hart). In order to

measure weight change of the material due to desorption of

water and solvent and decomposition of organics in TiO2

films, thermogravimetric analysis (TGA, TA instruments

2050) was performed in the presence of air at a ramp rate

of 3 -C/min.

2.4. Dye adsorption and photocatalytic activity

In order to measure dye adsorption of TiO2 films, the

prepared films were soaked into 3.0 mM methylene blue (MB,

Riedel-deHaen) solution, dried thoroughly, and analyzed using a

UV–Vis spectrophotometer (Hewlett Packard 8452A). During

the procedure, the initial blue color of MB-absorbed films was

changed to purple and the corresponding absorbance peak of

MB at 664 nm was shifted to around 564 nm. For photocatalytic

activity evaluation, the film with surface area of 10 cm2 was

placed into a borosilicate glass dish containing 8 ml of 30 AMMB solution at pH 3.0. Two 15 W low-pressure mercury UV

tubes (Spectronics) emitting near UV radiation with a peak at

365 nm were used at a light intensity of 3.48 mW/cm2 at the film

surface. The absorbance of MB solution at 664 nm was

monitored over time.

3. Results and discussion

3.1. Structural characteristics of TiO2 particles

In order to investigate the effect of surfactant type and

concentration on the structural properties of TiO2, porosimetry

and crystallographic analyses of TiO2 particles were employed,

and the results are summarized in Table 1. The structural

properties were significantly improved with the addition of

surfactant. In the absence of surfactant, the materials were

Table 1

Structural characteristics of TiO2 particles

Surfactant R SBET (m2/g) Vpore (cm3/g) Porosity (%) DBJH

Ads.

None 0.0 18.5 0.034 11.6 5.85

T20 0.5 37.0 0.054 17.4 4.82

1.0 50.7 0.071 21.6 4.50

2.0 98.0 0.257 50.0 8.52

2.5 104 0.309 54.6 9.68

3.0 109 0.372 59.1 11.5

T80 0.5 65.6 0.135 34.4 5.66

1.0 70.9 0.227 46.9 9.89

2.0 74.0 0.346 57.4 15.8

2.5 84.6 0.548 68.1 22.7

3.0 90.1 0.658 71.9 25.1

X100 0.5 60.5 0.098 27.6 5.16

1.0 85.6 0.144 35.9 5.35

2.0 85.8 0.148 36.5 6.08

2.5 87.5 0.240 48.3 9.16

3.0 87.3 0.238 48.1 11.3

a Calculated from BET surface area.b Estimated from XRD peak.c Measured from TEM image.

almost nonporous with surface area of 18.45 m2/g, pore

volume of 0.034 cm2/g, and porosity of 11.6%. Increasing R

from 0.0 to 1.0 resulted in 2.8–4.7 times increase in surface

area and 2.1–6.7 times increase in pore volume without

significant change in pore size. The structural characteristics

were consistent with the properties of the surfactants. It is well

known that increasing the surfactant chain length has a similar

effect as increasing surfactant concentration [23]. The higher

porosity and larger pore size of TiO2 materials prepared with

Tween surfactants compared to those prepared with X100 can

be ascribed to the fact that Tween surfactants have a chain of 20

ether groups while X100 bears a chain of only 10 ether groups.

Moreover, TiO2 prepared with T80 had the highest porosity

and largest pore size due to longer hydrophobic tail length of

T80 (C17) than T20 (C11). In addition, T80 had better

controllability of pore size in the mesoporous range.

The crystallite size (CSBET) was calculated from the BET

surface area, assuming all crystallites are spherical and separate

without aggregation [24]. The result was compared with the

crystallite size (CSXRD) estimated using Scherrer’s equation

from the XRD peak broadening analysis [25]. The CSBET was

always larger than the CSXRD, implying that most primary

particles were more or less aggregated. The actual crystallite

size (CSTEM) measured from TEM image slightly decreased

with increasing R, but it was within a similar range from 7 to

12 nm. In the case of R =0.0, the large discrepancy between

CSTEM and CSBET suggests the prepared particles were highly

aggregated. As a result, the addition of surfactants as pore

directing agents also inhibited the growth of crystallites as well

as the aggregation of adjacent primary particles, and thus made

the materials highly porous. The representative XRD patterns

of the TiO2 materials prepared with T80 are shown in Fig. 1.

All the peaks are designated to anatase crystal phase with size

less than 16 nm. Interestingly, more distinct peaks with higher

(nm) DBJH range (nm) CS, crystallite size (nm)

Des. Ads. Des. CSBETa CSXRD

b CSTEMc

5.24 3–11 2–9.1 83.4 15.6 11.9

4.36 2–11 2–8.5 41.5 10.8 9.87

4.28 2–10 2–8.1 30.3 10.0 8.85

7.35 4–23 4–18 15.7 9.45 7.65

8.12 6–29 6–22 14.7 8.25 7.42

9.45 8–30 8–23 14.2 8.25 7.45

5.23 3–12 3–8.9 20.6 12.4 9.65

9.24 4–18 4–15 20.8 12.0 9.23

13.2 8–36 8–22 20.0 10.7 8.89

18.6 14–40 14–24 18.2 10.6 9.34

20.1 17–48 17–28 17.1 11.2 9.54

4.26 3–11 3–8.0 25.4 12.0 10.7

4.65 3–12 3–8.4 18.0 10.5 9.05

5.04 3–14 3–8.9 16.6 12.0 8.53

7.62 5–32 5–15 17.6 12.0 8.89

8.64 7–33 7–20 17.5 12.3 9.32

Page 4: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

Molar ratio of surfactant to TTIP, R0.0 0.5 1.0 1.5 2.0 2.5 3.0

Abs

orba

nce

at 3

65 n

m (

arb.

uni

ts)

0.18

0.21

0.24

0.27

0.30

0.33

0.36

T80

T20X100

Fig. 3. UV light absorption of TiO2 films at 365 nm.20 30 40 50 60

Inte

nsity

(ar

b. u

nits

)

R=0.0

0.5

1.0

2.0

2.5

3.0

(101)(004) (200)

(105) (211)

2 (degrees)θ

Fig. 1. XRD patterns of TiO2 particles prepared with T80. Inserted numbers are

Miller indices.

H. Choi et al. / Thin Solid Films 510 (2006) 107–114110

intensities were observed with increasing R, suggesting the

material crystallinity was enhanced by the addition of

surfactants [26]. The same trend was observed in the case of

the other two surfactants.

3.2. Thickness, mass, UV light absorption, and dye adsorption

of TiO2 films

According to ESEM observation, the prepared transparent

TiO2 thin films were very reproducible and homogeneous

without cracks and pin-holes. As shown in Fig. 2, the thickness

of films increased as surfactant concentration R increased due

to an increase in the viscosity of the sol by the surfactants. The

mass of TiO2 immobilized on the glass substrates was

calculated from the thickness and porosity of films. In spite

of the increased film thickness, the mass of TiO2 decreased

over R because of the high porosity of the materials, which

implies that the TiO2 inorganic structure in the films became

less dense upon addition of surfactants.

Regardless of surfactant type and concentration, the UV

absorption edges of the TiO2 films were within the range of

372–374 nm, corresponding to band gap energy of around

3.34–3.32 eV. This is slightly larger than the commonly

reported band gap energy (Eg=3.23 eV, k =385 nm) for TiO2

Molar ratio of surfactant to TTIP, R0.0 0.5 1.0 1.5 2.0 2.5 3.0

Thi

ckne

ss (

μm)

0.24

0.26

0.28

0.30

0.32

0.34

0.36

Mas

s of

TiO

2 in

film

(μg

/cm

2 )

30

45

60

75

90

105

T80T20X100 T20

X100

T80

Fig. 2. Thickness and mass of TiO2 films.

Molar ratio of surfactant to TTIP, R0.0 1.0 1.5 2.0 2.5 3.0

Abs

orba

nce

at 5

64 n

m (

arb.

uni

ts)

0.0

0.2

0.4

0.6

0.8

1.0T80

T20

X100

0.5

Fig. 4. Visible light absorption of MB-adsorbed TiO2 films at 564 nm.

anatase phase because of the nanostructured properties of the

material [27,28]. As shown in Fig. 3, the UV light absorption

spectra of TiO2 films prepared under different conditions were

studied at 365 nm as an indirect evaluation of the photo-

catalytic activity under near UV radiation (300–400 nm).

Increasing surfactant concentration up to R =1.0 resulted in a

significant increase in the UV light absorption of TiO2 films,

and then increase in the UV light absorption became less

significant over R. This might be because the specific surface

area of the materials increased while the mass of TiO2

immobilized on the substrate decreased upon the addition of

surfactants. Moreover, the UV light absorption properties of

TiO2 film is a complex function of physicochemical properties

of the TiO2 material itself such as crystallinity, crystal phase,

crystal size, and purity as well as the structural properties of the

film such as thickness, mass, and surface area. In spite of their

smallest mass as shown in Fig. 2, filmsT80 had the highest UV

light absorption properties among films prepared with other

surfactants.

The visible light absorption of TiO2 films with adsorbed MB

at 564 nm as an indication of the porosity and homogeneity of

TiO2 films is shown in Fig. 4. MB adsorption of the films

increased with increasing surfactant concentration up to

R =1.0–2.0. In spite of the increased surface area of TiO2

material and thickness of the film with increasing surfactant

concentration, further addition of surfactant slightly decreased

Page 5: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

(a)

Relative Pressure (Ps/Po)

0.0 0.2 0.4 0.6 0.8 1.0

Vol

ume

Ads

orbe

d(c

m3 /g

ST

P)

0

20

40

60

80

100

120

140

160

R=0.0

R=1.0

Adsorption

Desorption

Por

evo

lum

e(c

m3 /

g)

R=1.0

R=0.0

(b)

Por

evo

lum

e(c

m3 /

g)

0.0

0.1

0.2

0.3

0.4

0.5

R=1.0

R=0.0

H. Choi et al. / Thin Solid Films 510 (2006) 107–114 111

the adsorption capacity due to a lower TiO2 mass immobilized

on the substrates. Another reason is perhaps associated with the

fact that the homogeneity of films prepared at too high

concentration of surfactants decreased significantly, most

probably, due to the high viscosity of the sol, pore coalescence,

and multi-micellar interactions.

3.3. Hydrophilicity of TiO2 films

Hydrophilicity (or wettability) of TiO2 films was investi-

gated by measuring the water contact angle. Surface hydro-

philicity is important for the accessibility of organic

compounds to catalytic sites [29]. It has been well observed

that the surface of TiO2 exhibits super-hydrophilicity under UV

irradiation due to hydrophilic groups introduced [30,31]. Prior

to coating with TiO2 films, the pre-cleaned glass substrate

showed water contact angle of around 16.1-, corresponding to

relatively hydrophilic surface. As shown in Fig. 5(a), the

contact angles after TiO2 coating significantly decreased to the

reliable detection limit of 4- (actually the contact angle reached

0-) without any UV radiation, indicative of super-hydrophilic

surface. The decreased contact angle may be associated with

the porous structure of TiO2 films and the increase in surface

hydrophilic groups, which was supported by TGA analysis of

TiO2 films. For the TGA analysis, approximately 10 mg of

particles were collected from the filmsT80 and the weight

change was monitored during thermal treatment at a ramp rate

(a)

Molar ratio of surfactant to TTIP, R

Molar ratio of surfactant to TTIP, R

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0

Con

tact

ang

le (

degr

ees)

4

6

8

10

12

14

16

T80X100T20

(b)

Wet

tabi

lity,

Cos

θ /

r

0.00

0.05

0.10

0.15

0.20

0.25

0.30

T80

X100

T20

Fig. 5. (a) Water contact angle and (b) wettability of TiO2 films. The contact

angle and pore size of TiO2 material are denoted as h and r, respectively.

Pore diameter (nm)

5 10 15 20 25 30

Pore diameter (nm)

5 10 15 20 25 30

Fig. 6. (a) Nitrogen adsorption–desorption isotherms and (b) pore size

distribution of TiO2 filmcontrol at R =0.0 and filmT80 at R =1.0. Inserted image

is side view of the filmT80 at R =1.0.

of 3 -C/min. The weight loss of the samples within the

temperature range 25–100 -C increased significantly from

1.98% for film prepared at R =0.0 to 2.31% for that at R =0.5.

Further increase in surfactant concentration up to R =3.0

caused slight increase in weight loss of 2.64%. This result

indicated that the TiO2 films prepared at high surfactant

concentration entrapped a large amount of water in the porous

structure.

In addition, according to the Laplace equation, the wet-

tability of porous materials is known to be proportional to the

surface tension of liquid and inversely proportional to the pore

size and surface energy (contact angle) of the material [32].

The increase in surfactant concentration resulted in decreased

contact angles as observed in Fig. 5(a) but increased pore size

as summarized in Table 1. Considering the same surface

tension of water used, Fig. 5(b) suggests that the optimum

surfactant concentration for preparing TiO2 films with better

hydrophilicity was at around R =0.5–1.0. FilmsT20 and

filmsX100 had a higher wettability than filmsT80 because more

organics were entrapped in films prepared with T80, which has

longer hydrocarbon chain length. According to TGA analysis

of films prepared with surfactants at R =1.0, negligible weigh

loss was observed after 100 -C in the case of filmT20 and

filmX100. On the other hand, filmT80 still had a considerable

Page 6: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

Table 2

Structural characteristics of TiO2 filmcontrol at R =0.0 and filmT80 at R =1.0

Parameter R =0.0 R =1.0

SBET (m2/g) 22.7 147

Vpore (cm3/g) 0.037 0.221

Porosity (%) 12.6 46.2

DBJH from adsorption branch (nm) 5.65 4.04

DBJH from desorption branch (nm) 5.38 3.72

CSTEM (nm) 12.4 9.20

Film thickness (Am) 0.26 0.31

TiO2 mass (Ag/cm2) 88.4 62.2

H. Choi et al. / Thin Solid Films 510 (2006) 107–114112

weight loss of 1.7% between 100 -C and 500 -C due to the

decomposition of organics remaining in the materials. More-

over, EDX results showed that the carbon content of filmT80

was around 1.2%, significantly higher than in other films,

which was in the range 0.4–0.6%.

3.4. Porosity and morphology of TiO2 films

Based on the obtained results so far, filmT80 at R =1.0

among all conditions investigated was considered the most

effective. For determining the structural properties of filmT80,

the TiO2 material was collected by scrapping the thin film

surface cautiously and analyzed using a porosimetry analyzer

and TEM. Fig. 6(a) shows the nitrogen adsorption–desorption

isotherms of filmT80. Compared to N2 isotherms of the

filmcontrol, which represents a nonporous material, these type

IV isotherms of filmT80 were typical of those of a well-

Fig. 7. Morphology and pore structure of (a–b) filmcontrol at R =

developed mesoporous material. A hysteresis loop in the

isotherms was observed with dissimilar shapes for the

adsorption and desorption branches, implying a different size

of pore throat diameter. The sharp drop on the desorption

branch can be assigned to the presence of mesopore constric-

tions at the boundaries between the ordered domains and of

smaller pores in the titania walls [33]. The pore size

distribution shown in Fig. 6(b) was relatively narrow ranging

from 2 to 8 nm. The Barrett, Joyner and Halenda (BJH) pore

diameters measured from the adsorption and desorption

branches were 4.04 nm and 3.72 nm, respectively. These

results imply good homogeneity of the pores. The main

structural characteristics deduced from the isotherms are

reported in Table 2. In spite of the high heat treatment

temperature of 500 -C, the BET surface of 147 m2/g and

porosity of 46.2% were significantly high, compared to other

research results reported [5,17]. The film thickness of 0.31 Amwas measured using ESEM, and its mass of 62.2 Ag/cm2 was

calculated from the pore volume and density of the anatase

crystal phase. Even though the thickness of filmT80 was much

larger than that of filmcontrol, the amount of TiO2 catalyst in

filmT80 was smaller due to its high porosity.

As shown in Fig. 6(b), the cross-section of filmT80 on glass

substrate revealed that filmT80 was homogeneous and well-

incorporated to the glass substrates without cracks and pin-

holes. EDX elemental analysis of collected thin films showed

that the films were composed of mainly Ti and O elements

without any significant impurities. Fig. 7 shows the morphol-

0.0 and (c–d) filmT80 at R =1.0 at different magnifications.

Page 7: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

Reaction time (h)

Nor

mal

ized

MB

abs

orba

nce,

I/I o

0.0

0.2

0.4

0.6

0.8

1.0Without TiO2

R=0.0

R=1.0R=2.0

R=3.0

R=0.5

1 2 3 4 5 6 70

Fig. 8. Photocatalytic decoloration of MB by TiO2 filmsT80.

H. Choi et al. / Thin Solid Films 510 (2006) 107–114 113

ogy of the nanostructured anatase TiO2 thin films. For

filmcontrol, no distinct mesopore structure was observed and

even the lattice fringes were not clear. On the other hand,

filmsurfactant were highly porous and exhibited distinct pore

structure. The films had slightly collapsed spherical bicontin-

uous structure with highly interconnected network [34]. The

image at high magnification showed many randomly oriented

nanocrystallites with size of 8.3–10.6 nm and sets of clearly

resolved lattice fringes giving evidence that the TiO2 material

was highly crystalline, which was in good agreement with the

XRD results. The porous structure was very strong and stable

since the structure still remained at large extent until heat

treatment at 700 -C.

3.5. Photocatalytic activity of TiO2 films

Photocatalytic activity of the TiO2 films was measured in

terms of MB decoloration. As shown in Fig. 8, there was no

direct photolysis of MB in the absence of TiO2 photocata-

lysts. In spite of its relatively high catalyst mass of 88.4 Ag/cm2, the filmcontrol at R =0.0 did not show significant

decoloration of MB due to the almost nonporous properties

of the film, suggesting the photocatalytic reaction occurred

only at the very surface of TiO2 film. Increasing the amount

of surfactant template up to R =1.0–2.0 resulted in a

significant improvement of the photocatalytic activity of the

films mainly due to the porous structure of the film with high

surface area of 147 m2/g and porosity of 46%, and partially

the enhanced material crystallinity as observed in the XRD

and TEM analyses. However, further addition of surfactant

beyond R =2.0 caused adverse effect on the photocatalytic

activity of films, which was consistent with the results on MB

adsorption. It was because the homogeneity of the films

prepared under too high concentration of surfactants (R >2.0)

decreased and the amount of TiO2 photocatalyst immobilized

TiO2 films also decreased. These results show the importance

of preparing highly porous and homogeneous TiO2 thin films,

which might facilitate MB adsorption and UV light utiliza-

tion. Moreover, considering the small amount of TiO2 catalyst

immobilized on the glass substrate, the TiO2 films were

highly efficient to decolorize the dye.

4. Conclusions

Highly porous and hydrophilic nanostructured TiO2 thin

films and particles were synthesized from nanocomposite

organic/inorganic sol–gel, composed of isopropanol, acetic

acid, titanium tetraisopropoxide, and nonionic surfactant

molecules as templates. Slow hydrolysis reaction and stable

incorporation of the inorganic network onto surfactant

molecules made it possible to control the subsequent porous

TiO2 nanostructure. The TiO2 particles and films had

enhanced structural and catalytic properties including high

surface area, large pore volume, pore size controllability, small

crystallite size, enhanced crystallinity, and active anatase

phase. Among the surfactants investigated, Tween 80 was

the most promising in terms of film homogeneity, UV

absorbance, methylene blue adsorption, and especially pore

volume and pore size controllability. The porous TiO2 thin

films were highly efficient to decolorize MB dyes due to their

high surface area and photocatalytic activity. This acetic acid-

based sol–gel method modified by varying the type and the

concentration of the surfactant template is useful in the

preparation of nanostructured anatase TiO2 thin films with

high photocatalytic activity and desired pore structure for

environmental applications.

Acknowledgement

This research was funded by a grant from the Office of

Biological and Physical Research of the National Aeronautics

and Space Administration (NRA Grant No. NAG 9-01475).

References

[1] M. Koelsch, S. Cassaignon, C. Ta Thanh Minh, J.-F. Guillemoles, J.-P.

Jolivet, Thin Solid Films 451 (2004) 86.

[2] K. Yoo, H. Choi, D.D. Dionysiou, Chem. Commun. (2004) 2000.

[3] M. Langlet, A. Kim, M. Audier, C. Guillard, J.M. Herrmann, Thin Solid

Films 429 (2003) 13.

[4] G. Balasubramanian, D.D. Dionysiou, M.T. Suidan, V. Subramanian, I.

Baudin, J.M. Laıne, J. Mater. Sci. 38 (2003) 823.

[5] F. Bosc, A. Ayral, P.-A. Albouy, C. Guizard, Chem. Mater. 15 (2003)

2463.

[6] D.D. Dionysiou, A.A. Burbano, M.T. Suidan, I. Baudin, J.M. Laıne,

Environ. Sci. Technol. 36 (2002) 3834.

[7] S.-Y. Kwak, S.H. Kim, S.S. Kim, Environ. Sci. Technol. 35 (2001) 2388.

[8] D.M. Antonelli, Microporous Mesoporous Mater. 30 (1999) 315.

[9] P. Yang, D. Zhao, D.I. Margolese, B.F. Chmelka, G.D. Stucky, Chem.

Mater. 11 (1999) 2813.

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

[11] N. Idrissi-Kandri, A. Ayral, M. Klotz, P.-A. Albouny, A.E. Mansouri, A.

Van der Lee, C. Guizard, Mater. Lett. 50 (2001) 57.

[12] D. Trong On, Langmuir 15 (1999) 8561.

[13] S. Barboux-Doeuff, C. Sanchez, Mater. Res. Bull. 29 (1994) 1.

[14] C. Sanchez, J. Livage, M. Henry, F. Babonneau, J. Non-Cryst. Solids 100

(1998) 65.

[15] P. Dunbar, Birnie III, N.J. Bendzko, Mater. Chem. Phys. 59 (1999) 26.

[16] C. Wang, Z.X. Deng, Y. Li, Inorg. Chem. 40 (2001) 5210.

[17] E. Stathatos, P. Lianos, C. Tsakiroglou, Microporous Mesoporous Mater.

75 (2004) 255.

[18] O. Dag, I. Soten, O. Celik, S. Polarz, N. Coombs, G.A. Ozin, Adv. Funct.

Mater. 13 (2003) 30.

Page 8: Synthesis of nanocrystalline photocatalytic TiO2 thin films and particles using sol–gel method modified with nonionic surfactants

H. Choi et al. / Thin Solid Films 510 (2006) 107–114114

[19] E. Prouzet, F. Cot, G. Nabias, A. Larbot, P. Kooyman, T.J. Pinnavaia,

Chem. Mater. 11 (1999) 1498.

[20] M. Klotz, N. Idrissi-Kandri, A. Ayral, C. Guizard, Mater. Res. Soc. Symp.

Proc. (2000) 628.

[21] R.S.A. de Lange, J.H.A. Hekkink, K. Keizer, A.J. Burggraaf, J. Non-

Cryst. Solids 195 (1996) 203.

[22] C.-Y. Tsai, S.-Y. Tam, Y. Lu, C.J. Brinker, J. Membr. Sci. 169 (2000) 255.

[23] J. Zhang, Z.-L. Wang, J. Liu, S. Chen, G.-Y. Liu, Self-Assembled

Nanostructure, Kluwer Academic/Plenum Publishers, New York, 2003.

[24] M. Ivanda, S. Music, S. Popovic, M. Gotic, J. Mol. Struct. 480 (1999)

645.

[25] S. Nakade, M. Matsuda, S. Kambe, Y. Saito, T. Kitamura, T. Sakata, Y.

Wada, H. Mori, S. Yanagida, J. Phys. Chem., B 106 (2002) 10004.

[26] R.J. Young, Introduction to Polymers, Chapman and Hall, New York,

1981.

[27] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, J. Phys. Chem. 92

(1988) 5196.

[28] A. Henglein, Ber. Bunsenges, Phys. Chem. 86 (1982) 301.

[29] S. Hata, Y. Kai, I. Yamanaka, H. Oosaki, K. Hiroto, S. Yamagishi, JSAE

Rev. 21 (2000) 97.

[30] T. Wanatabe, A. Nakajima, R. Wang, M. Minabe, S. Koizumi, A.

Fujishima, K. Hashimoto, Thin Solids Films 351 (1999) 260.

[31] H. Irie, S. Washizuka, N. Yoshino, K. Hashimoto, Chem. Commun.

(2003) 1298.

[32] M. Mulder, Basic Principles of Membrane Technology, Kluwer Academic

Publishers, Netherlands, 1991.

[33] P.I. Ravikovitch, A.V. Neimark, Langmuir 18 (2002) 1550.

[34] P.A.A. Alberius, K.L. Frindell, R.C. Hayward, E.J. Kramer, G.D. Stucky,

B.F. Chmelka, Chem. Mater. 14 (2002) 3284.