hydrogen peroxide-assisted photocatalytic oxidation of phenolic compounds

6
Hydrogen peroxide-assisted photocatalytic oxidation of phenolic compounds M.A. Barakat a, * , J.M. Tseng b , C.P. Huang b a Central Metallurgical R&D Institute, Helwan 11421,Cairo, Egypt b Department of Civil and Environmental Engineering, University of Delaware, DE 19716, USA Received 17 September 2004; received in revised form 10 January 2005; accepted 14 January 2005 Available online 5 February 2005 Abstract The effect of hydrogen peroxide (H 2 O 2 ) on photocatalytic oxidation of phenol and monochlorophenols (CP) in aqueous suspensions of commercial TiO 2 rutile was investigated. Various concentrations of H 2 O 2 were used without and with the presence of TiO 2 under different atmospheres, e.g., N 2 or O 2 . Sources of hydroxyl radicals for photocatalytic processes are suggested through the surface hydroxyl group reacting with hole, dissolved oxygen trapping an electron, and photolytic H 2 O 2 . The combination of TiO 2 and H 2 O 2 under UV illumination can greatly enhance the degradation rates of the phenol and chlorophenols. The photocatalytic oxidation with the H 2 O 2 /UV/TiO 2 system was found to be much more effective than either UV/TiO 2 or UV/H 2 O 2 alone. The efficiency of the photocatalytic degradation of phenol was improved from 30 to 97% due to the presence of H 2 O 2 . As the H 2 O 2 concentration increases, more hydroxyl radicals are produced, and the phenol oxidation rate increases. At high H 2 O 2 concentration (10 2 M), O 2 or N 2 atmospheres are not important factors for phenol oxidation in the H 2 O 2 /UV/TiO 2 system. # 2005 Elsevier B.V. All rights reserved. Keywords: H 2 O 2 /UV/TiO 2 system; Photocatalytic degradation; H 2 O 2 concentration 1. Introduction Phenol and phenolic substances are used as raw materials in many petrochemical, chemical, and pharmaceutical industries. Wastewater containing phenol has received increased attention because of its toxicity and prevalence in industrial processes [1–3]. In addition, phenol is considered to be an intermediate product in the oxidation of higher-molecular weight aromatic hydrocarbons [4]. Thus, it is usually taken as a model compound for advanced wastewater treatment studies. Different methods were reported for phenol degradation in aqueous solutions. Catalytic wet oxidation is one of the conventional methods which uses H 2 O 2 as oxidative agent with homogenized metal salts [5,6], or a combination of H 2 O 2 with heterogeneous catalysts such as Al–Fe pillared clay or iron containing clays [7–9]. Advanced oxidation by using UV/ H 2 O 2 is another way to degrade phenolic compounds in dilute solutions [10–15]. The UV/H 2 O 2 processes result in formation of OH radicals, which accelerate the oxidative degradation of the phenolic compounds in water. In recent years, photocatalytic degradation of phenolic compounds in the presence of semiconductor powders such as titanium dioxide (TiO 2 ) has been considered as effective technology [16–25]. TiO 2 photocatalyst upon irradiation with UV light produces electrons and positive holes. The holes are strong oxidizing agents that can oxidize organic compounds such as phenols to mineral acids, e.g., HCl and CO 2 . The main disadvantage of using H 2 O 2 is its high cost, whereas the low oxidation rate is the major draw back of TiO 2 . The combined use of H 2 O 2 and TiO 2 is a way of solving this problem. Fewer studies have been devoted to the examination of phenol and chlorophenols degradation in presence of H 2 O 2 and TiO 2 anatase [26–29]. In this study commercial TiO 2 rutile was treated and investigated as a photocatalyst. The effect of H 2 O 2 on improving the reactivity of the catalyst for phenol and www.elsevier.com/locate/apcatb Applied Catalysis B: Environmental 59 (2005) 99–104 * Corresponding author. Tel.: +20 25010642; fax: +20 25010639. E-mail address: [email protected] (M.A. Barakat). 0926-3373/$ – see front matter # 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.apcatb.2005.01.004

Upload: ma-barakat

Post on 04-Sep-2016

218 views

Category:

Documents


3 download

TRANSCRIPT

Hydrogen peroxide-assisted photocatalytic oxidation of

phenolic compounds

M.A. Barakat a,*, J.M. Tseng b, C.P. Huang b

a Central Metallurgical R&D Institute, Helwan 11421,Cairo, Egyptb Department of Civil and Environmental Engineering, University of Delaware, DE 19716, USA

Received 17 September 2004; received in revised form 10 January 2005; accepted 14 January 2005

Available online 5 February 2005

Abstract

The effect of hydrogen peroxide (H2O2) on photocatalytic oxidation of phenol and monochlorophenols (CP) in aqueous suspensions of

commercial TiO2 rutile was investigated. Various concentrations of H2O2 were used without and with the presence of TiO2 under different

atmospheres, e.g., N2 or O2. Sources of hydroxyl radicals for photocatalytic processes are suggested through the surface hydroxyl group

reacting with hole, dissolved oxygen trapping an electron, and photolytic H2O2. The combination of TiO2 and H2O2 under UV illumination

can greatly enhance the degradation rates of the phenol and chlorophenols. The photocatalytic oxidation with the H2O2/UV/TiO2 system was

found to be much more effective than either UV/TiO2 or UV/H2O2 alone. The efficiency of the photocatalytic degradation of phenol was

improved from 30 to 97% due to the presence of H2O2. As the H2O2 concentration increases, more hydroxyl radicals are produced, and the

phenol oxidation rate increases. At high H2O2 concentration (�10�2 M), O2 or N2 atmospheres are not important factors for phenol oxidation

in the H2O2/UV/TiO2 system.

# 2005 Elsevier B.V. All rights reserved.

Keywords: H2O2/UV/TiO2 system; Photocatalytic degradation; H2O2 concentration

www.elsevier.com/locate/apcatb

Applied Catalysis B: Environmental 59 (2005) 99–104

1. Introduction

Phenol and phenolic substances are used as raw materials

in many petrochemical, chemical, and pharmaceutical

industries. Wastewater containing phenol has received

increased attention because of its toxicity and prevalence

in industrial processes [1–3]. In addition, phenol is

considered to be an intermediate product in the oxidation

of higher-molecular weight aromatic hydrocarbons [4].

Thus, it is usually taken as a model compound for advanced

wastewater treatment studies. Different methods were

reported for phenol degradation in aqueous solutions.

Catalytic wet oxidation is one of the conventional methods

which uses H2O2 as oxidative agent with homogenized

metal salts [5,6], or a combination of H2O2 with

heterogeneous catalysts such as Al–Fe pillared clay or iron

containing clays [7–9]. Advanced oxidation by using UV/

* Corresponding author. Tel.: +20 25010642; fax: +20 25010639.

E-mail address: [email protected] (M.A. Barakat).

0926-3373/$ – see front matter # 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.apcatb.2005.01.004

H2O2 is another way to degrade phenolic compounds in

dilute solutions [10–15]. The UV/H2O2 processes result in

formation of OH radicals, which accelerate the oxidative

degradation of the phenolic compounds in water. In recent

years, photocatalytic degradation of phenolic compounds in

the presence of semiconductor powders such as titanium

dioxide (TiO2) has been considered as effective technology

[16–25]. TiO2 photocatalyst upon irradiation with UV light

produces electrons and positive holes. The holes are strong

oxidizing agents that can oxidize organic compounds such

as phenols to mineral acids, e.g., HCl and CO2. The main

disadvantage of using H2O2 is its high cost, whereas the low

oxidation rate is the major draw back of TiO2. The combined

use of H2O2 and TiO2 is a way of solving this problem.

Fewer studies have been devoted to the examination of

phenol and chlorophenols degradation in presence of H2O2

and TiO2 anatase [26–29].

In this study commercial TiO2 rutile was treated and

investigated as a photocatalyst. The effect of H2O2 on

improving the reactivity of the catalyst for phenol and

M.A. Barakat et al. / Applied Catalysis B: Environmental 59 (2005) 99–104100

monochlorophenols (CP) photocatalytic degradation was

studied. Two systems, H2O2 alone and H2O2 with TiO2,

under different atmospheres (O2 or N2) were investigated.

Table 1

Properties of the used titanium dioxide (rutile [17] and Degussa P25)

Property DuPont TiO2 Degussa P25 TiO2

Raw Treated

Structure Rutile Rutile Anatase:rutile (80:20)

pHzpc 9.3 9.0 6.6

Surface area (m2/g) 6.6 6.1 50 � 1.5

Band gab energy (eV) 3.2 3.2 3.2

Impurities (wt.%) 0.729 (Al) 0.708 (Al) 0.1

1.145 (Cl) 0.0 (Cl) 0.0

2. Methods and materials

2.1. Materials

Titanium dioxide (TiO2), commercial grade, was

provided by the DuPont Company (Wilmington, Delaware,

USA). Raw TiO2 powder was treated according to Tseng and

Huang [17] by rinsing with 1 M HClO4 solution followed by

washing with distilled water. Degussa P25 TiO2 was

provided by Degussa Company (Ridge-Field Park, NJ,

USA) and used without any pretreatment. Stock solutions of

both Dupont and Degussa TiO2 (10 g/L) were prepared,

shaken for 24 h and left for 3 days for hydration. The TiO2

suspension was shaken for 30 min every time before use.

Hydrogen peroxide was prepared from a 34% H2O2 stock

solution and diluted to the required concentration. All

phenols that used in this study were obtained from Aldrich

Chemical Co. (98–99%) and were used without further

purification.

The light sources used to illuminate TiO2 samples was a

1600 W medium pressure mercury vapor discharge lamp

(American Ultraviolet Co.). The spectral irradiance for the

UV lamp (260 W/m2) ranges from 228 to 420 nm at a

distance of 1 m from the light source according to the

information provided by the manufacturer. The light

intensity was recorded with a UV-radiometer (Model

365H, Spectronics Co.). A light intensity of 17.0 W/m2,

at a peak of 365 nm was provided at 1 m distance from test

tube reactors. The light wavelength (greater than 325 nm)

can completely pass through the Pyrex glass [17].

2.2. Methods

Pyrex glass tubes having dimensions of 20 mm internal

diameter and 13 cm long were used as reactors. All photo-

catalytic oxidation experiments were performed in the Pyrex

tubes containing 15 mL of 10�3 M phenol solution and

0.15 g of TiO2. A constant ionic strength (I = 5 � 10�2 M)

was obtained by addition of a given amount of NaClO4 to the

solutions. In order to study the effect of oxygen, different

atmospheres were created by introducing pure nitrogen or

oxygen into the reactors. Head space of each reactor also

was filled with the same gas as in solution. A pyrogallol

solution was used to remove oxygen prior to nitrogen

bubbling. The reactors were tightened with Teflon septa

airtight caps. The phenol solutions were aerated with

nitrogen or oxygen prior to experiments. Samples were

shaken over a reciprocal shaker (American Optical Co.) with

180 strokes per minute to insure complete mixing. The

temperature of the solutions was controlled and adjusted to

25 8C by a thermostat pump in conjunction with a cooler. At

the end of a given reaction time, samples were filtered using

0.45 mm microfilter (Gelman, Supor-450, 25 mm diameter).

2.3. Measurements

Dissolved oxygen was monitored by a dissolved oxygen

(DO) meter calibrated by the modified Winkler method [30].

Generally, 0.3 ppm oxygen was found in solution under

nitrogen bubbling. In the pure oxygen atmosphere 27.0 ppm

dissolved oxygen can be obtained. The residual concentra-

tion of phenol was measured with an UV–vis spectro-

photometer (Hitachi/Perkin-Elmer, Model 139) at a

wavelength of 271 nm. The optimal wavelength was

determined from the spectrum of phenol solutions at various

pHs and concentrations. Chloride was measured by using an

Orion Model 96-176 ion specific combination chloride

electrode.

3. Results and discussion

The syntheses of the Dupont TiO2 catalyst were already

optimized and characterized in a previous study [17]. All

photo oxidation experiments were performed with the

pretreated Dupont TiO2, one run of photocatalytic degrada-

tion of phenol was carried out with Degussa P25 TiO2 for

comparison with the data of Dupont TiO2. The major

properties of the used TiO2, Dupont [17] and Degussa P25

(data supplied by the manufacturer, are summarized in

Table 1. The photo oxidation experiments for both phenol

and chlorophenols were conducted at optimized experi-

mental conditions that were reported in the previous works

[17,18]. The experimental conditions were: UV light

intensity of 17 W/m2, solution temperature of 25 8C,

solution pH of 7, and phenol or CP concentration of

10�3 M. The ionic strength of the solutions was adjusted by

adding 5 � 10�2 M NaClO4. At such experimental condi-

tions with dilute phenol concentration, it was very difficult

to identify intermediates of the degradation. The inter-

mediates were detected by GC/MS, and the data was

reported in a previous work [17]. Two intermediates,

hydroquinone and phenol dimmer were detected only at

higher phenol concentration, 5 � 10�2 M. Fig. 1 shows the

effect of H2O2 on the photolytic phenol oxidation under N2

M.A. Barakat et al. / Applied Catalysis B: Environmental 59 (2005) 99–104 101

Fig. 1. Effect of H2O2 on photolytic oxidation of phenol under N2 or O2

atmosphere. Experimental conditions: 10�3 M phenol, hn = 17 W/m2,

pH = 7, I = 5 � 10�2 M, temperature = 25 8C, H2O2 = 10�2 M.

Fig. 3. Effect of H2O2 concentration on photocatalytic oxidation of phenol

with TiO2 in O2 system. Experimental conditions: 10�3 M phenol,

hn = 17 W/m2, pH = 7, I = 5 � 10�2 M, temperature = 25 8C, TiO2 =

10 g/L.

or O2 atmosphere. In the absence of the H2O2, phenol was

not oxidized at all either in N2 or in O2 atmosphere. By using

10�2 M H2O2 only, phenol oxidation in an O2 saturated

solution was low (20% removal after 6 h), while there was

no oxidation in N2 solution. Photolysis of H2O2 produces

hydroxyl radicals as follows:

H2O2�!hv

2�OH (1)

wherein, the molar extinction coefficient of H2O2 at 254 nm

is 19.6 M�1 s�1 [31]. The applied wavelength in this work

was 365 nm at which H2O2 has an extremely low absorption

[32]. Therefore the oxidation of phenol by photolytic H2O2

was insignificant. Fig. 2 presents the photolytic phenol

degradation under O2 in the presence of different concen-

trations of H2O2. Phenol degradation has not been found to

be significant at the early time up to 2 h (incubation time),

the degradation rate increased with increase in the H2O2

Fig. 2. Effect of H2O2 concentration on photolytic oxidation of phenol in O2

system. Experimental conditions: 10�3 M phenol, hn = 17 W/m2, pH = 7,

I = 5 � 10�2 M, temperature = 25 8C.

concentration. The degradation efficiency reached a

maximum value of 55% after 6 h with 10�1 M H2O2. In

homogenous catalysis, the reaction rate is determined by the

Brownian-motion collision of reactants [33]. The higher the

concentration, the better the reaction rate. Phenol oxidation

in UV/H2O2 is low in a homogeneous system. Although

hydroxyl radicals are claimed to exist in such system, the

reaction rate is relatively small due to the probability of

collision as follows:

H2O2 þ �OH!HO2� þ H2O (2)

H2O2 þ HO2� ! �OH þ H2O þ O2 (3)

Fig. 3 demonstrates the phenol photocatalytic oxidation at

various H2O2 concentrations in an O2 atmosphere over TiO2.

In the UV/TiO2 system (10 g/L TiO2, no H2O2), the rate of

phenol photodegradation increased gradually with time. The

efficiency of the photodegradation was 30% after 2 h and it

reached a maximum efficiency value of 94% after 6 h.

According to photocatalytic processes, the electrons and

holes are separated by photoexcitation of semiconductors

(e.g., TiO2). The holes migrate to the surface and react with

adsorbed species while the electrons react with O2 to form

hydroxyl radicals. It has been known that O2 plays an

important role in trapping electrons in this process. The

rate of phenol photodegradation increased significantly in

the H2O2/UV/TiO2 system. It seems that 5 � 10�3 M H2O2

was enough to occupy the surface sites of TiO2. Consistent

results of phenol degradation were obtained with different

concentrations of H2O2. The degradation rate increased with

increase in the H2O2 concentration reaching a maximum

value of 97% after 2 h with 10�2 M H2O2.

Fig. 4 demonstrates the phenol photocatalytic oxidation

at various H2O2 concentrations in an N2 atmosphere over

TiO2. In the absence of H2O2, the lack of O2 resulted in

increase in the surface recombination rate. Therefore, the

phenol oxidation reaction appeared to be slow in the N2

M.A. Barakat et al. / Applied Catalysis B: Environmental 59 (2005) 99–104102

Fig. 4. Effect of H2O2 concentration on photocatalytic oxidation of phenol

with TiO2 in N2 system. Experimental conditions: 10�3 M phenol,

hn = 17 W/m2, pH = 7, I = 5 � 10�2 M, temperature = 25 8C, TiO2 =

10 g/L.Fig. 5. Effect of TiO2 on photocatalytic oxidation of phenol with H2O2 in

O2 system. Experimental conditions: 10�3 M phenol, hn = 17 W/m2,

pH = 7, I = 5 � 10�2 M, temperature = 25 8C, H2O2 = 10�2 M.

system. Upon adding H2O2, the rate of phenol photode-

gradation increased significantly by the same manner and at

the same conditions with O2 atmosphere. At high H2O2

concentration (�10�2 M), O2 and N2 were not important

factors for phenol oxidation at the applied experimental

conditions. The presence of H2O2 can enhance the rate of

photocatalytic process. This can be explained as follows; the

recombination of holes (h+) and electrons (e�) has been

regarded as an unfavorable or limiting process in photo-

catalysis. The dissolved O2 reduces the effect of the

recombination of charges and acts as scavenger for the

photon-produced electrons on TiO2 surface [29], according

to the following reaction:

e�ðTiO2Þ þ O2 !�O�2 (4)

With H2O2 the overall photocatalytic oxidation rate can be

enhanced via direct photolysis of H2O2 by UV light with the

production of hydroxyl radicals (Eq. (1)) which are likely to

be the dominant rate-improving mechanism in this process

[32]. Also in the absence of O2, H2O2 can compensate for the

O2 lack and play a role as an external electron scavenger

according to the following reactions:

e�ðTiO2Þ þ H2O2 !�OH þ OH� (5)

H2O þ hþ!�OH þ Hþ (6)

Fig. 6. Rate of initial oxidation of phenol vs. TiO2 concentration.

A chain reaction of H2O2 decomposition can be triggered

after the initiation of the hydroxyl radicals [34]. In the

absence of H2O2, �OH via the generated H2O2 are negligible

compared with that with holes. This is due to the difficulty in

reducing oxygen to form H2O2 at the illuminated TiO2. On

contrary, in the presence of H2O2, the �OH generation is a

direct result. The hydroxyl radical formation, either via

photo excitation of water molecules or H2O2 adsorption

onto solid, relies on the availability of the hydrated TiO2

surface. Different amounts of TiO2 were added to 10�2 M

H2O2 and the results are presented in Fig. 5. In the absence of

TiO2, hydroxyl radicals that were formed have difficulty

reacting with phenol. In addition, TiO2 improved the contact

opportunity between hydroxyl radicals and phenol, thereby

greatly increased the phenol degradation rate. It can be seen

that 1 g/L of TiO2 was enough to degrade 97% of the phenol

after 2 h irradiation. Increasing TiO2 concentration up to 5 g/

L resulted in almost no change in the degradation efficiency,

however further increase in TiO2 dosage decreased the

phenol degradation efficiency. The relationship between

the initial rate of phenol oxidation and TiO2 concentration

is shown in Fig. 6. Results indicate that the overall reaction

was controlled by surface reactions, i.e. the active sites and

the photo-absorption of the catalyst used. Adequate loading

of the catalyst increased the generation rate of electron–hole

pairs for promoting the degradation of phenol. However, at

high dose of the TiO2, the light penetration by the photo-

M.A. Barakat et al. / Applied Catalysis B: Environmental 59 (2005) 99–104 103

Fig. 7. Efficiency of phenol photodegradation with the used catalysts (TiO2

rutile and Degussa P25) in the absence and in the presence of H2O2.

Experimental conditions: 10�3 M phenol, hn = 17 W/m2, pH = 7,

I = 5 � 10�2 M, temperature = 25 8C, H2O2 = 10�2 M.

Fig. 8. Photolytic oxidation of monochlorophenols with H2O2. Experi-

mental conditions: 10�3 M phenol, hn = 17 W/m2, pH = 7, I = 5 � 10�2 M,

temperature = 25 8C, O2 atmosphere, H2O2 = 10�2 M.

Fig. 9. Photocatalytic oxidation of monochlorophenols in the presence of

H2O2. Experimental conditions: 10�3 M phenol, hn = 17 W/m2, pH = 7,

I = 5 � 10�2 M, temperature = 25 8C, O2 atmosphere, H2O2 = 10�2 M,

TiO2 = 10 g/L.

catalyst suspension decreased due to an increase in turbidity,

thereby the reaction rate decreased [35]. To realize the

improving effect of the H2O2 on the photocatalytic process,

two runs of phenol degradation were performed by using

Degussa P25 TiO2 in the absence and in the presence of

H2O2 at the same experimental conditions that used with the

TiO2 rutile. Fig. 7 shows the efficiency of phenol photo-

degradation with both of TiO2 rutile and Degussa P25 in the

absence and in the presence of H2O2 as a function of the

irradiation time under the same experimental conditions

(2.5 g/L catalyst, O2 atmosphere, 10�2 M H2O2, 10�3 M

phenol, and pH 7). In the absence of H2O2, the degradation

efficiency of phenol increased gradually with time reaching

maximum values of 94 and 99.5% after 6 h of irradiation

with the rutile and the Degussa P25, respectively. It is worthy

to note that the rate of degradation with Degussa P25 was

faster than that with rutile. The rate of the degradation

efficiency was improved with both catalysts in the presence

of H2O2. After 2 h of irradiation, the efficiency value of

degradation was 97% with H2O2/UV/TiO2 (rutile) compared

to 30% with UV/TiO2 (rutile), while a complete degradation

was achieved with the H2O2/UV/Degussa P25 system as

compared with 88% with UV/Degussa P25 system. This

confirms the improving effect of the H2O2 on both of the rate

and the efficiency of phenol oxidation which can be attrib-

uted to the mentioned oxidant species (Eqs. (5) and (6)).

Direct photolysis (H2O2/UV system) on monochloro-

phenol oxidation was investigated. Fig. 8 presents the results

of monochlorophenol oxidation with 10�2 M H2O2 at

different chloride positions: ortho- (2-CP), meta- (3-CP),

and para-position (4-CP). The results show that the para- and

meta-positions were favored for chlorophenol photodegra-

dation. Complete degradation was achieved at 3 h with 3-CP

and 4-CP, while with 2-CP the photodegradation efficiency

was about 75%. As reported previously [17], the mechanism

for photocatalytic oxidation of phenol suggests that the OH

radical attacks the C4 (para) position of phenol as a result of

hydroquinone formation. It seems that a similar mechanism

can be applied to the chlorophenol oxidation.

The effect of H2O2 on the photocatalytic oxidation of

monochlorophenol was also investigated. Fig. 9 presents the

results of H2O2/UV/TiO2 system with monochlorophenol.

Complete degradation was achieved after 30 min with 3-CP

and 4-CP, while with 2-CP it was achieved after 2 h. The

addition of H2O2 increased the concentration of hydroxyl

radicals on the TiO2 surface, therefore, the attack on

chlorophenol was more flexible. The chloride attached to the

M.A. Barakat et al. / Applied Catalysis B: Environmental 59 (2005) 99–104104

Fig. 10. Chloride production by photocatalytic oxidation of monochlor-

ophenols in the presence of H2O2. Experimental conditions: 10�3 M phenol,

hn = 17 W/m2, pH = 7, I = 5 � 10�2 M, temperature = 25 8C, O2 atmo-

sphere, H2O2 = 10�2 M, TiO2 = 10 g/L.

benzene ring can be replaced by OH radicals simultaneously

without position preference. According to the results of

chloride production (Fig. 10), the reaction is expected to be

complete and less intermediates are expected to be formed.

The photocatalytic oxidation of chlorophenols proceeds

much faster than chloride release, this is due to the formation

of chlorinated intermediates [20]. A comparison of

photolytic and photocatalytic results confirms that the

oxidation reaction is most likely happening on the surface of

TiO2. Generally, chlorophenol has better adsorption than

phenol onto TiO2.

4. Conclusions

The combined use of H2O2 and UV light can greatly

enhance the performance of TiO2 as a photocatalyst. The

photocatalytic oxidation with the H2O2/UV/TiO2 system

was found to be much more effective than either UV/TiO2 or

UV/H2O2 alone. H2O2 produces hydroxyl radicals under UV

illumination, which increase the rate of photocatalytic

oxidation of phenol and chlorophenols. The organic

pollutants are attacked by both UV photons and hydroxyl

radicals generated from H2O2. As the H2O2 concentration

increases, more hydroxyl radicals are produced, and the

phenol oxidation rate increases. The efficiency of the

photocatalytic degradation of phenol was improved from 30

to 97% due to the H2O2. At high H2O2 concentration

(�10�2 M), O2 or N2 atmospheres are not important factors

for phenol oxidation in the H2O2/UV/TiO2 system.

References

[1] J.W. Moore, S. Ramamoorthy, Organic Chemicals in Natural Waters,

Springer-Verlag, New York, 1984.

[2] M.D. Mann, W.G. Wilson, J.G. Gasifier, Environ. Prog. 4 (1) (1985)

33.

[3] I.M.M. Yurii, S. Moshe, Ind. Eng. Chem. Res. 37 (1998) 309.

[4] S. Aurora, Y. Pedro, D. Beatriz, G.O. Felix, Ind. Eng. Chem. Res. 40

(2001) 2773.

[5] F.J. Rivas, S.T. Kolaczkowski, F.J. Beltran, D.B. McLurgh, J. Chem.

Technol. Biotechnol. 74 (5) (1999) 390.

[6] H. Grigoropoulou, C. Philippopoulos, Water Sci. Technol. 36 (2/3)

(1997) 151.

[7] K. Fajerwerg, T. Castan, A. Paerrard, H. Debellefontaine (Eds.), in:

Proceedings of the International Conference on Oxidation Technol-

ogies for Water and Wastewater TreatmentGoslar, Germany, May 12–

15, 1996, p. 8.

[8] J. Guo, M. Al-Dahhan, Ind. Eng. Chem. Res. 42 (2003) 2450.

[9] E. Guelou, J. Barrault, J. Fournier, J.M. Tatibouet, Appl. Catal. B:

Environ. 44 (2003) 1.

[10] J. Chen, W.H. Rulkens, H. Bruning, Water Sci. Technol. 35 (4) (1997)

231.

[11] A.K. De, B. Chaudhuri, S. Bhattacharjee, B.K. Dutta, J. Hazard. Mater.

64 (1999) 91.

[12] R. Alnaizy, A. Akgerman, Adv. Environ. Res. 4 (2000) 233.

[13] M. Hugul, R. Apak, S. Demirci, J. Hazard. Mater. V (2000) 193.

[14] M.Y. Ghaly, G. Hartel, R. Mayer, R. Haseneder, Waste Manage. 21

(2001) 41.

[15] D.H. Han, S.Y. Cha, H.Y. Yang, Water Res. 38 (2004) 2782.

[16] R.W. Matthews, Water Res. 24 (5) (1990) 653.

[17] J.M. Tseng, C.P. Huang, in: D.W. Tedder, F.G. Pohland (Eds.),

Emerging Technologies in Hazardous Waste Management, ACS

Symposium Series 422, ACS, Washington, DC, 1990p. 12.

[18] J.M. Tseng, C.P. Huang, Water Sci. Technol. 23 (1–3) (1991) 377.

[19] U. Stafford, K.A. Gray, P.V. Kamat, J. Phys. Chem. 98 (1994) 6343.

[20] C. Dong, C.P. Huang, in: C.P. Hugul, C.O. O’Melia, J.J. Morgan

(Eds.), Aquatic Chemistry: Interfacial and Interspecies Processes,

Advances in Chemistry Series 244, ACS, Washington, DC, 1995p.

291.

[21] M. Hugul, I. Boz, R. Apak, J. Hazard. Mater. 64 (1999) 313.

[22] D. Chen, A.K. Ray, Appl. Catal. B: Environ. 23 (1999) 143.

[23] L. Rideh, A. Wehrer, D. Ronze, A. Zoulalian, Catal. Today 48 (1999)

357.

[24] H. Chun, W. Yizhong, T. Hongxiao, Chemosphere 41 (2000) 1205.

[25] A.M. Peiro, J.A. Ayllon, J. Peral, X. Domenech, Appl. Catal. B:

Environ. 30 (2001) 359.

[26] T. Wei, Y. Wang, C. Wan, J. Photochem. Photobiol. A: Chem. 55

(1990) 115.

[27] R.M. Alberici, W.F. Jardim, Water Res. 28 (8) (1994) 1845.

[28] R. Bauer, G. Waldner, H. Fallmann, S. Hager, M. Klare, T. Krutzler, S.

Malato, P. Maletzky, Catal. Today 53 (1999) 131.

[29] S.G. Poulopoulos, C.J. Philippopoulos, J. Environ. Sci. Health A:

Toxic/Hazard. Subst. Environ. Eng. A39 (6) (2004) 1385.

[30] Standard Methods, 421 B, Azide Modification, 15th ed., 1980, p. 390.

[31] Y. Ogata, K. Tomizawa, Y. Yamashita, J. Chem. Soc., Perkin Trans. II

(1980) 616.

[32] W. Chu, C.C. Wong, Water Res. 38 (2004) 1037.

[33] A. Morita, Bussei Kenkyu 54 (1990) 147.

[34] T.Y. Wei, Y.Y. Wang, C.C. Wan, Photochem. Photobiol. A: Chem. 55

(1990) 115.

[35] A. Doong, W.H. Chang, J. Photochem. Photobiol. A: Chem. 116

(1998) 221.