Você está na página 1de 7

Applied Catalysis A: General 304 (2006) 55–61

www.elsevier.com/locate/apcata

Photocatalytic degradation of organic contaminants


in water by ZnO nanoparticles: Revisited
C. Hariharan *
Organic Chemistry Division, National Chemical Laboratory, Pune, India
Received 4 October 2005; received in revised form 30 January 2006; accepted 7 February 2006
Available online 24 March 2006

Abstract
Nanoscale photocatalysts have attracted much attention due to their high surface area to volume ratios. This work investigates the
photodegradation of organic contaminants using the fluorescence emission characteristics of ZnO nanoparticles (ZnO-nano) in aqueous solutions.
This is accomplished by preparing nanocrystalline ZnO; the presence of organic contaminants in water is readily detected from the quenching of
fluorescence observed from ZnO semiconductor films. Photolysis of ZnO thin films immersed into an aqueous system containing organic
contaminants results in the degradation of the contaminants. A comprehensive study has been done involving several organic contaminants in water
(like aliphatic and aromatic chloro compounds as well as some commonly used aromatic solvents) to check the suitability of ZnO-nano as an
efficient photocatalyst. The ZnO nanoparticles not only serve as a better catalytic system compared to bulk ZnO and commercially available
Degussa TiO2 in achieving degradation of the added contaminants, but unlike other semiconductor systems can also act as a non-specific sensor for
the presence of these common contaminants in water. A total cleanup of a cocktail of contaminants in water was also achieved using the ZnO-nano.
# 2006 Elsevier B.V. All rights reserved.

Keywords: ZnO; Nanoparticles; Semiconductor; Organic contaminants and photocatalysis

1. Introduction In recent years, semiconductor photocatalysis has become


more and more attractive and important since it has a great
Contaminations of ground water systems by organic potential to contribute to such environmental problems. One
chemicals pose a serious environmental threat. The severity of the most important aspects of environmental photocatalysis
of this threat is due to their toxicity to animals and humans. The is in the selection of semiconductor materials like ZnO and
‘‘Environmental Protection Agency’’ (USA) has listed several TiO2, which are close to being two of the ideal photocatalysts
volatile organic compounds as contaminants in ground water. in several respects. For example, they are relatively
The list includes several chlorinated aromatic and aliphatic inexpensive, and they provide photo-generated holes with
compounds like 4-chlorophenol, pentachlorophenol, chloro- high oxidizing power due to their wide band gap energy. As a
benzene, dichlorobenzene, carbon tetrachloride, chloroform, well-known photocatalyst, ZnO has received much attention
dichloroethylene, trichloroethylene and dichloromethane, and in the degradation and complete mineralization of environ-
other common organic solvents like phenols, toluene, benzene mental pollutants [6–9]. Since ZnO has almost the same band
and xylenes [1]. All possible clean-up routes like microbial gap energy (3.2 eV) as TiO2, its photocatalytic capacity is
degradation, direct photodegradation or even hydrolysis are anticipated to be similar to that of TiO2. However, in the case
painfully slow on these compounds [2]. Several groups have of ZnO, photocorrosion frequently occurs with the illumina-
investigated photocatalytic oxidation of organic compounds to tion of UV light, and this phenomenon is considered as one of
break them to simpler fragments or have attempted a complete the main reasons for the decrease of ZnO photocatalytic
mineralization using different metal oxide-semiconductor activity in aqueous solutions [10,11]. However, some studies
materials [3–5]. have confirmed that ZnO exhibits a better efficiency than TiO2
in photocatalytic degradation of some dyes, even in aqueous
solution [12,13]. Kormann compared the photocatalytic
* Tel.: +91 20 25902076; fax: +91 20 25893153. activity values of ZnO, Fe2O3 and TiO2; the results indicated
E-mail addresses: c.manikandan@ncl.res.in, chithramani@yahoo.com. that ZnO and TiO2 exhibited higher activity than Fe2O3 during
0926-860X/$ – see front matter # 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2006.02.020
56 C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61

the process of photocatalytic oxidation of gas phase (12.6 g) was dissolved in ethanol (200 mL) at 50 8C. The
chloridized hydrocarbon [14]. Although ZnO and TiO2 in oxalic acid solution was added slowly under conditions
general have been proved to be very active in the photo- of stirring to the warm ethanolic solution of zinc acetate.
catalytic oxidation of different pollutants, the catalyst A thick white gel formed, which was kept for drying at
morphology is very crucial in determining the performance 80 8C for 20 h. The xerogel was calcined at different
efficiency of the catalytic system, and also in deciding about temperatures (400, 500, 600 and 800 8C) for 2 h to yield ZnO
the lifetimes and the deactivation and regeneration of the nanoparticles [16]. UV–vis, photoluminescence, IR spe-
semiconductor material which are very crucial in environ- ctroscopy, X-ray diffraction (XRD), atomic force microscopy
mental photocatalysis. (AFM) and BET surface area analysis were used for
Nanometer-sized particles have very different physical characterization.
and chemical properties from bulk materials. When used as
catalysts, their catalytic activity is expected to be enhanced 2.2. Casting of ZnO films
not only because of their increased surface area, but also
because of the changes of surface properties such as surface One milligram of ZnO powder was suspended in 1 mL of
defects. When the crystallite dimension of a semiconductor double distilled water and sonicated for 15 min. The ZnO
particle falls below a critical radius of approximately 50 nm, suspension was used for casting films on plain glass or quartz
the charge carriers appear to behave quantum mechanically plates. A small aliquot (usually 100–150 mL) was spread over
like simple particles in a box. This confinement produces a an area of 1 cm2 and air-dried. The semiconductor coated glass
quantization of discrete electronic states and increases the plates were then annealed in the oven at 673 K for 1 h. The
effective band gap of the semiconductor [15]. As a result, the immobilized semiconductor films were used for quenching
band edges shift to yield larger redox potentials. The solvent experiments and for photodegradation studies.
reorganization free energy for charge transfer to a substrate,
however, remains unchanged. The increased driving force 2.3. UV–vis and IR spectroscopy
resulting from the increased redox potential and the
unchanged solvent reorganization free energy in size- Absorption spectra were recorded using a UV–vis spectro-
quantized systems are expected to increase the rate constant photometer Hitachi-Model 330. The optical path length of the
of charge transfer and may result in increased photo cuvettes used in all the experiments was 1 cm. The absorption
efficiencies for systems in which the rate-limiting step is spectra were recorded at room temperature (23–25 8C). The
charge transfer. In the present work, I tested the suitability of absorption spectra of ZnO-nano suspension was taken in water.
ZnO nanoparticles as a better photocatalyst in degrading Infrared spectra were recorded as pressed KBr-discs on a
common organic contaminants as compared to bulk ZnO and Shimadzu FT-IR-8300 spectrophotometer.
commercial TiO2-degussa (P25). ZnO nanoparticles were
prepared via a non-aqueous route and the effect of different 2.4. Photoluminescence spectroscopy
calcination temperatures on surface area as well as on the
photocatalytic activity of ZnO was studied. Results indicate A Perkin-Elmer LS-50B spectrofluorimeter was used for
that in general the photocatalytic activity of ZnO-nano was steady-state fluorescence measurements. For recording of the
superior to those of bulk-ZnO as well as TiO2 (Degussa). The fluorescence spectra, unpolarized light from the Xenon lamp
efficiency of the catalytic activity showed dependence on was used. As far as possible, narrow excitation and emission slit
particle size, crystallinity and surface area. A complete widths were chosen for both the excitation and the emission
cleanup of a synthetic organic cocktail in water was sides. Excitation was at 325 nm. For studies with ZnO films,
effectively achieved using the ZnO-nano. Hence, ZnO-nano front face illumination was used.
by itself or mixed with other semiconductor oxides (like
TiO2) can serve as a good photocatalyst to identify and 2.5. Atomic force microscopy
clean up a variety of common organic contaminants found
in water. A VEECO digital Instruments multimode scanning probe
microscope equipped with a nanoscope-IV controller was
2. Materials and methods operated in contact mode to characterize the surface
morphology of the ZnO films. The average particle size for
Zinc acetate dihydrate, oxalic acid dihydrate and all the ZnO was determined from AFM measurements.
organic reagents were procured from Merck, India. Double
distilled water was used for all the measurements. 2.6. X-ray diffraction

2.1. Preparation of ZnO nanoparticles A Rigaku Geigerflex X-ray diffractometer with Ni-filtered
Cu Ka radiation (l = 1.5406 Å, 40 kV, 30 mA) was used to
In a typical experiment, zinc acetate dihydrate (10.98 g) determine the crystallinity and phase purity of the samples.
was treated with ethanol (300 mL) at 60 8C. The salt was The XRD patterns were recorded in the 2u range of 20–708,
completely dissolved in about 30 min. Oxalic acid dihydrate with a scan speed of 28/min.
C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61 57

Fig. 1. IR spectrum of ZnO-nano (500 8C).


Fig. 2. X-ray powder diffraction plot of ZnO-bulk and ZnO-nano (calcined at
500 8C).

2.7. BET surface area analysis 3. Results and discussion

The specific surface area (SBET) was estimated from the N2 3.1. Characterization of ZnO-nano
adsorption/desorption isotherms, measured on an Omnisorb
100CX Coulter instrument. IR analysis of the dried and calcined (at 500 8C) ZnO
powder shows characteristic bands at 487 and 517 cm 1 [17].
2.8. Photolysis There is a small shoulder around 3420 cm 1 which could be
surface adsorption of moisture. A commercially available
All the photocatalysis experiments were carried out in a ZnO powder sample (data not shown) also shows surface OH
closed photocell arrangement, which has 15 mL of water groups around 3400 cm 1. Similar features have been observed
having 50 mM of organic contaminant and ZnO thin films for ZnO nanoparticles/colloids obtained by aqueous precipita-
coated glass/quartz slide was immersed into the setup. The tion. A comparison with our samples (Fig. 1) indicates that a
solution was kept under constant stirring. The ZnO film was non-aqueous route may intrinsically lower the surface coverage
illuminated continuously with light from a 450 W Henovoniea with hydroxylic groups [17]. The particle size and morphology
mercury lamp filtered through a solution of CuSO4. Samples of the obtained ZnO were characterized by powder X-ray
were drawn periodically from the reservoir for the UV-analysis. diffraction (Fig. 2) and by AFM (Fig. 3). The powder X-ray

Fig. 3. AFM image of ZnO-nano film coated on a glass slide.


58 C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61

diffraction pattern shows 2u values and relative intensities of chlorinated compounds, we recorded emission spectra of ZnO
the peaks that coincide with the JCPDS data of zincite. The thin films immersed into an aqueous solution containing
average crystallite size for ZnO-nano calcined at 500 8C was different concentrations of the organic contaminants (20–
determined as 35  5 nm from the linewidth broadening of the 250 mM). The solution was continuously stirred during the
XRD peak corresponding to (0 0 2) reflection, using the emission measurements to assist contact between ZnO
Debye–Scherrer equation. The corresponding AFM image nanoparticles and the organic contaminants. Fig. 5 shows
(Fig. 3) shows spherical ZnO nanoparticles, with fairly uniform
size. The mean diameter of these particles is roughly 39 nm.
The photoluminescence spectra were taken for the aqueous
suspensions and the ZnO films (ZnO-nano immobilized in a
quartz slide) (Fig. 4). The excitonic absorption band was
observed at 325 nm, blue shifted due to a quantum confinement
effect compared to that of the bulk ZnO (usually seen at
385 nm). The photoluminescence spectra (lex = 325 nm) of the
ZnO nanoparticles show a strong UV emission at 390 nm;
three weak emitting bands (blue emission at 430 nm, a blue–
green band at 480 nm and a green band at 520 nm) were
observed. The green band emission corresponds to the singly
ionized oxygen vacancy in ZnO and results from the
recombination of a photogenerated hole with the single ionized
charge state of this defect. The weak green emission also
implies that there are few surface defects in the ZnO
nanoparticles. The strong UV emission should correspond to
the near band edge emission of the wide band gap of ZnO, due
to the annihilation of excitons and a result of the quantum
confinement effect. The excellent room temperature UV
emission property should be attributed to the high purity and
perfect crystallinity of the synthesized ZnO nanoparticles [18].

3.2. Fluorescence quenching of ZnO-nano by contaminants

To check to see whether ZnO emission is sensitive to the


presence of common contaminants like aromatic solvents and

Fig. 5. Extent of quench as a function of contaminant concentration (excitation


wavelength set at 325 nm) fo is the emission intensity of the ZnO-nano in the
absence of quencher and f is the emission intensity in the presence of the
quencher. The amount of quench increases as a function of quencher concen-
tration. Panel A: (*) benzene, (&) toluene and (~) ortho-xylene. Panel B: (*)
4-chlorophenol, (&) chlorobenzene and (~) 1,4-dichlorobenzene. Panel.C:
Fig. 4. Room-temperature photoluminescence and absorption (inset) spectrum (*) carbon tetrachloride, (&) chloroform, (~) dichloromethane, (!) methy-
of ZnO-nano. lene chloride, (*) dichloroethane and (*) tetrachloroethane.
C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61 59

the drop in fluorescence intensity for the immobilized ZnO Table 1


Apparent association constant for adsorption of contaminants onto ZnO-nano
thin films as a function of the organics present in aqueous
films
solution. The fluorescence emission intensity decreased with
increasing concentration of the contaminant. The origin for Ka (M 1)
fluorescence in the ZnO-nano films is due to a charge Solvents
recombination process. As the organic compound from Benzene 6235
aqueous solution came in contact with the semiconductor Toluene 6878
o-Xylene 7234
surface, it effectively scavenged the holes, thus competing
with the charge recombination process responsible for ZnO Aromatic compounds
emission [19]. The organic compounds taken up for study can Chlorophenol 13858
Chlorobenzene 9678
be grouped into three different categories (Fig. 5). The Dichlorobenzene 7435
chloroaromatics show the largest extent of quench, followed
Aliphatic compounds
by the aromatic solvents (benzene, toluene and xylene).
Carbon tetrachloride 3829
The aliphatic chlorocompounds were comparatively weak Chloroform 3678
quenchers. Dichloromethane 2123
A pre-requisite for such a charge transfer at the Methylene chloride 2240
semiconductor interface is effective adsorption of organic Dichloroethane 2819
Trichloroethane 2620
molecules. The surface concentration of the adsorbed species
Tetrachloroethane 2514
and the energetics determine the rate and efficiency of hole
transfer at the semiconductor interface. We have analyzed the
emission quenching data by considering adsorption equili- quenching effect was expected for the aliphatic chlorocom-
brium between ZnO and the organic contaminants [20]. One pounds, because the presence of aromatic rings as well as the
representative sample in each group is shown in Fig. 6. A number and nature of substituents on the ring (like electron
straight-line fit of the double reciprocal plot for the drop in donating or electron withdrawing groups) are known to affect
fluorescence as a function of contaminant concentration, adsorption [20].
shown in Fig. 6, confirms a strong adsorption of the organic
contaminants onto ZnO-nano. From the intercept and slope of 3.3. Photocatalytic degradation of contaminants
the straight-line plot, an apparent equilibrium constant was
obtained (Table 1). The quenching effects seen amongst the To test for the photodegradation of organic contaminants
three groups in Fig. 5 (chloroaromatics > aromatics > by ZnO-nano, I introduced ZnO immobilized thin films into
chloroaliphatic) were paralleled by the adsorption results, in the solution containing organic contaminants and photoirra-
that the weakest quenchers (aliphatic chloro compounds) also diated them using a mercury lamp. At periodic intervals of time,
showed weak adsorption onto the ZnO surface. A weak aliquots of the sample were withdrawn and the absorption
spectra were recorded. The absorption band around 280 nm
for the aromatic contaminants decreased as a function of
time. This can be attributed as oxidative degradation of the
organic compound by ZnO. Direct bandgap excitation of the
semiconductor results in electron–hole separation. Photogen-
erated holes oxidize the organic contaminant at the ZnO
surface.
The decrease in the absorbance seen at 280 nm for the
aromatic compounds as a function of time can be used to estimate
the percentage degradation of the contaminant (Fig. 7). After
3–4 h of irradiation, maximum loss of the aromatic species is
noted, indicating that the reaction has attained steady-state. No
noticeable change occurs at longer durations. However,
photoirradiation of the contaminants without ZnO-nano resulted
in no change in the absorption spectra. The quenching efficiency
(shown in Fig. 5) and the percentage degradation of organic
contaminants (shown in Fig. 7) are dictated by the effective
adsorption of the organic contaminants onto the ZnO-nano (Ka
values listed in Table. 1). The surface concentration of the
adsorbed species will dictate the efficiency of charge transfer
at the semiconductor interface responsible for the catalytic
activity. The binding isotherm (shown in Figs. 5 and 6) depicts
Fig. 6. Plot of I0/(I0 I) vs. 1 /[contaminant]. Ka is obtained from slope/ ‘‘specific binding’’ characteristics and saturate at higher
intercept [20]. (*) 4-Chlorophenol, (&) benzene and (~) CCl4. contaminant concentration.
60 C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61

Fig. 7. Percentage degradation of the contaminant as a function of time of


photoirradiation. (*) 4-Chlorophenol, (&) chlorobenzene, (~) 1,4-dicholro- Fig. 8. Performance comparision of different catalytic systems. The rate of
benzene and (!) benzene. degradation of 4-chlorophenol by different catalytic systems is shown.

3.4. Ability of ZnO-nano to sense cocktails 3.7. The catalytic activity and stability of the ZnO-nano

The ability of ZnO-nano to sense mixtures of contaminants This was tested for atleast four repeated cycles. During the
in water was also checked using a synthetic mixture of several course of the photoreaction, the reaction mixture was alternated
aromatic compounds (like chlorophenols and chloroben- between 0.5 mM of organic contaminant and deionized water
zenes) in water. The photoirradiation experiments were and the emission response of ZnO films was tested (data not
repeated on the mixture. After 4 h of irradiation, almost 85% shown). The following was observed:
loss of the aromatic species was observed (data not shown).
This indicates that ZnO-nano can simultaneously sense and (i) ZnO emission is restored once the reaction vessel is flushed
degrade mixtures of contaminants in water. with deionized water. This confirms that ZnO is reactivated
just by flushing in water.
3.5. Comparison of catalytic activities (ii) Adsorption between the organic contaminant and the ZnO
is chemisorption in nature.
The photocatalytic activities of commercial TiO2 and bulk (iii) The emission traces showed that the activity of ZnO is not
ZnO were compared with that of the ZnO-nano. The initial rates lost even after several repeated switchings between the
per unit mass of catalyst for the decomposition of 4- organic contaminant and water.
chlorophenol degradation by different catalyst are plotted in
Fig. 8. The efficiency of ZnO-nano as a catalyst is superior 3.8. Effect of pH
compared to those of both micron-sized bulk ZnO and
commercial TiO2 (Degussa-P25). The pH of the reaction mixture was altered to study the
photodegradation taking place on semiconductor particle
3.6. Effect of calcination temperature on ZnO morphology surfaces, since pH can alter the surface charge properties of

The effect of varying the calcination temperature of the Table 2


xerogel to yield ZnO-nano also had an effect on the particle Comparision of different catalysts
size and BET surface area which affected the efficiency of the Catalyst Calcination Average BET surface
ZnO-nano (as seen from Table 2 and Fig. 8). Higher temperature (8C) particle size area (m2 g 1)
calcination temperatures resulted in increases in the particle ZnO 400 42  5 nm 19
size and a drop in surface area; accompanied by a drop in ZnO-nano 500 35  5 nm 27
catalytic activity; this suggests that, in addition to surface ZnO 600 48  5 nm 15
area, the morphology of ZnO-nano (the particle size and ZNO 800 62  5 nm 9
ZnO-Bulk 500 0.5 mm 5
crystallinity) also plays a crucial role in deciding the catalyst
TiO2 (Degussa P25) Commercial 25–30 nm 40–50
performance.
C. Hariharan / Applied Catalysis A: General 304 (2006) 55–61 61

the photocatalyst. Effect of pH on the photodegradation of References


4-CC was investigated over the pH range of 4–12 (data not
shown). A steady increase in the degradation of 4-CC is [1] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95
(1995) 69.
observed up to 9.0 followed by a slow decrease at higher pH
[2] J.P. Wilcoxon, J. Phys. Chem. 104 (2000) 7334.
value. Higher pH value can provide higher concentration of [3] R.H. Mills, D. Davies, Worsley, Chem. Soc. Rev. (1993) 417.
hydroxyl ions to react with holes (h+) to form hydroxyl radicals [4] N. Serpone, P. Muthukumar, P. Pickat, E. Pelizzetti, H. Hidaka, J.
(OH), subsequently enhancing the photodegradation rate of Photochem. Photobiol. A: Chem. 85 (1995) 247.
4-CC. But at pH values higher than 10, photocatalytic activity [5] R.W. Mathews, J. Phys. Chem. 91 (1987) 3328.
of ZnO is strongly influenced by pH value, since its stability is [6] C. Richard, F. Bosquet, J.F. Pilichowski, J. Photochem. Photobiol. A:
Chem. 108 (1997) 45.
not as good as other semiconductors. [7] M.D. Driessen, T.M. Miller, V.H. Grassian, J. Mol. Catal. A: Chem. 131
(1998) 149.
4. Conclusions [8] J. Villase_oor, P. Reyes, G.J. Pecchi, Chem. Technol. Biotechnol. 72
(1998) 105.
The suitability of ZnO nanoparticles (both as suspension [9] M.C. Yeber, J. Rodrıguez, J. Freer, N. Durian, H.D. Mansilla, Chemo-
sphere 41 (2000) 1193.
and thin films) to sense and photodegrade several common [10] V. Dijken, A.H. Janssen, M.H.P. Smitsmans, D. Vanmaekelbergh, K.
organic contaminants were tested. ZnO-nano was prepared Meijerink, Chem. Mater. 10 (1998) 3513.
using a non-aqueous route and different experimental [11] S. Neppolian, B. Sakthivel, M. Arabindoo, V. Palanichamy, J. Murugesan,
conditions have been used to tailor the performance of Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 34 (9)
(1999) 1829.
ZnO-nano as a photocatalyst. The results show that a wide
[12] K. Gouvea, F. Wypych, S.G. Moraes, N. Duran, N. Nagata, P. Peralta-
variety of aromatic and aliphatic chlorocompounds listed as Zamora, Chemosphere 40 (2000) 433.
contaminants present in water can quench the fluorescence [13] S. Dindar, J. Icli, Photochem. Photobiol. A: Chem. 140 (2001) 263.
from ZnO-nano; hence this system has been used to sense the [14] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, J. Photochem. Photo-
presence of contaminants and to photodegrade them effec- biol. A 48 (1989) 161.
tively. This study confirms that use of ZnO nanoparticles as a [15] K.Y. Jung, Y.C. Kang, S.B. Park, J. Mater. Sci. Lett. 22 (1997) 1848.
[16] S.C. Pillai, J.M. Kelly, D.E. McCormack, P. O’Brien, R. Ramesh, J. Mater.
catalytic system is comparable to and sometimes even better Chem. 13 (2003) 2586.
than the conventional photocatalyst of choice-TiO2-Degussa [17] Y.J. Kwon, K.H. Kim, C.S. Lim, k.B. Shim, J. Cer. Proc. Res. 3 (2002) 146.
(P25). Future research should examine the effectiveness of [18] J. Wang, L. Gao, J. Mater. Chem. 13 (2003) 2551.
this catalyst in degrading real industrial effluents, in which [19] P.V. Kamat, B. Patrick, J. Phys. Chem. 96 (1992) 6829.
numerous organic contaminants are present. [20] P.V. Kamat, R. Huehn, R. Nicolaescu, J. Phys. Chem. 96 (2002) 788.

Você também pode gostar