Você está na página 1de 7

Photocatalytic degradation of amoxicillin, ampicillin and cloxacillin antibiotics in

aqueous solution using UV/TiO


2
and UV/H
2
O
2
/TiO
2
photocatalysis
Emad S. Elmolla , Malay Chaudhuri
Department of Civil Engineering, Universiti Teknologi PETRONAS, 31750 Tronoh, Perak, Malaysia
a b s t r a c t a r t i c l e i n f o
Article history:
Received 10 September 2009
Received in revised form 4 November 2009
Accepted 4 November 2009
Available online 27 November 2009
Keywords:
Amoxicillin
Ampicillin
Cloxacillin
Antibiotics
UV/H
2
O
2
/TiO
2
photocatalysis
Degradation of amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution by TiO
2
photocatalysis
under UVA (365 nm) irradiation was studied. Enhancement of photocatalysis by addition of H
2
O
2
was also
evaluated. The results showed that no signicant degradation occurred by 300-min UVA irradiation per se
and pH had a great effect on antibiotic degradation. Photocatalytic reactions approximately followed a
pseudo-rst order kinetics and the rate constants (k) were 0.007, 0.003 and 0.029 min
1
for amoxicillin,
ampicillin and cloxacillin, respectively. Addition of H
2
O
2
at ambient pH5 and TiO
2
1.0 g/L resulted in
complete degradation of amoxicillin, ampicillin and cloxacillin in 30 min. Dissolved organic carbon (DOC)
removal, and nitrate (NO
3

), ammonia (NH
3
) and sulphate (SO
4
2
) formation during degradation indicated
mineralization of organic carbon, nitrogen and sulphur. UV/H
2
O
2
/TiO
2
photocatalysis is effective for
degradation of amoxicillin, ampicillin and cloxacillin in aqueous solution.
2009 Elsevier B.V. All rights reserved.
1. Introduction
Pharmaceutical compounds including antibiotics have been
observed in the aqueous environment. These compounds have been
observed in surface water [13], ground water [3], sewage efuent
[4,5] and even in drinking water [6]. Pharmaceutical compounds can
reach the aquatic environment through various sources such as
pharmaceutical industry, hospital efuent and excretion from
humans and livestock [5,7,8]. Among all the pharmaceutical com-
pounds that cause contamination of the environment, antibiotics
occupy an important place due to their high consumption rate in both
veterinary and human medicine. A problem that may be created by
the presence of antibiotics in low concentration in the environment is
the development of antibiotic resistant bacteria [9]. In recent years,
the incidence of antibiotic resistant bacteria has increased and many
people believe that the increase is due to the use of antibiotics [10].
Furthermore, the presence of antibiotics in wastewaters has also
increased and their abatement will be a challenge in the near future.
It is useful to apply various treatment technologies to purify
efuents containing pharmaceutical compounds. The advanced
oxidation processes (AOPs) appear as interesting tools in comparison
with other techniques such as activated carbon adsorption, air
stripping and reverse osmosis. Indeed, many of these techniques
only transfer the pollutants from one phase to another without
destroying them. Biological treatment is limited to wastewaters
which contain biodegradable substances and which are not toxic to
the biological culture. Among the different advanced oxidation
processes, TiO
2
photocatalysis has emerged as a promising wastewa-
ter treatment technology. The main advantages of the process are lack
of mass transfer limitations and operation at ambient conditions, and
the catalyst is inexpensive, commercially available, non-toxic and
photochemically stable [11].
Reaction mechanisms of photocatalytic processes have been
discussed in the literature [1217]. Illumination of a semiconductor
such as titanium dioxide (TiO
2
) by photons having an energy level
that exceeds its band gap energy (hvNE
bg
=3.2 eV in the case of
anatase TiO
2
) excites electrons (e

) from the valence band to the


conduction band and holes (h
+
) are produced in the valence band
(Reaction (1)). The photogenerated valence band holes react with
either water (H
2
O) or hydroxyl ions (OH

) adsorbed on the catalyst


surface to generate hydroxyl radicals (
U
OH) which are strong oxidants
(Reactions (2) and (3)). The hydroxyl radical reacts readily with
surface adsorbed organic molecules, either by electron or hydrogen
atom abstraction, forming organic radical cations, or by addition
reactions to unsaturated bonds [17] (Reaction (4)). Since the reaction
of the holes on the particle interface is faster than electrons, the
particles under illumination contain an excess of electrons. Removal
of these excess electrons is necessary to complete the oxidation
reaction by preventing the recombination of electrons with holes. The
most easily available electron acceptor is molecular oxygen and in the
presence of oxygen the predominant reaction of electrons is that with
O
2
to formradicalic superoxide ions (
U
O
2

) as in Reaction (5). In acidic


condition, superoxide ion combines with proton to form a hydroper-
oxide radical and it reacts with conduction band electron to form
Desalination 252 (2010) 4652
Corresponding author. Tel.: +60 14 9047313.
E-mail address: em_civil@yahoo.com (E.S. Elmolla).
0011-9164/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2009.11.003
Contents lists available at ScienceDirect
Desalination
j our nal homepage: www. el sevi er. com/ l ocat e/ desal
hydroperoxide ion. The hydroperoxide ion reacts with proton to form
hydrogen peroxide. Cleavage of hydrogen peroxide by the conduction
band electrons yields further hydroxyl radicals and hydroxyl ions
(Reaction (6)). The hydroxyl ions can then react with the valence
band holes to formadditional hydroxyl radicals. Recombination of the
photogenerated electrons and holes may occur and indeed it has been
suggested that preadsorption of substrate (organic substance) onto
the photocatalyst is a prerequisite for highly efcient degradation.
TiO
2
hv TiO
2
e

1
h

H
2
OH


U
OH 2
h

OH

U
OH 3
Organics
U
OHDegradation Products 4
e

O
2

U
O

2
5
H
2
O
2
e

U
OH OH

6
The photocatalytic processes have been reported to be effective for
degradation of many organic pollutants such as carboxylic acids
[18,19], chloroanilines [20], mono-chlorocarboxylic acids [21], chlor-
ophenols [2224], dyes [2528], fungicides [29], herbicides [30],
ketones [31], phenolics [32] and pharmaceuticals [8,33]. Degradation
of lincomycin and sulfamethoxazole antibiotics by TiO
2
photocatalysis
have been reported [33,34].
Amoxicillin, ampicillin and cloxacillin are semi-synthetic penicillin
obtaining their antimicrobial properties fromthe presence of a -lactam
ring. They are widely used in human and veterinary medicine. Some
authors have found amoxicillin and cloxacillin in wastewater [35,36].
There are some reported studies on degradation of amoxicillin by
different AOPs. Zhang et al. [37] studied the treatment of amoxicillin
wastewater by combination of extraction, Fenton oxidation and
reverse osmosis. Also, different AOPs (O
3
/OH

, H
2
O
2
/UV, Fe
2+
/H
2
O
2
,
Fe
3+
/H
2
O
2
, Fe
2+
/H
2
O
2
/UV and Fe
3+/
H
2
O
2
/UV) were evaluated as a
pretreatment process for real penicillin formulation wastewater [38].
The degradation of amoxicillin by ozonation [39] and photo-Fenton
[40] has been reported. In our previous work, we studied the deg-
radation of amoxicillin, ampicillin and cloxacillin antibiotics using
Fenton [41], photo-Fenton [42] and ZnO photocatalysis [43].
However, no study on degradation of amoxicillin, ampicillin and
cloxacillin antibiotics in aqueous solution by TiO
2
photocatalysis has
been reported.
The present study was undertaken to examine the degradation of
amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution
by UV/TiO
2
photocatalysis and the enhancement of photocatalysis by
hydrogen peroxide addition.
2. Materials and methods
2.1. Chemicals
Analytical grades of amoxicillin (AMX) and ampicillin (AMP) were
purchased fromSigma and cloxacillin (CLX) fromFluka and they were
used to construct HPLC analytical curves for determination of
antibiotic concentration. AMX, AMP and CLX used to prepare
antibiotic aqueous solution were obtained from a commercial source
(Farmaniage Company). The commercial products were used as
received without any further purication. Sodium hydroxide and
sulphuric acid were purchased from HACH Company USA. Potassium
dihydrogen phosphate (KH
2
PO
4
) was purchased from Fluka and
acentonitrile HPLC grade from Sigma. TiO
2
powder (anatase, purity
N99%) was purchased from Fluka. Hydrogen peroxide (30% w/w) was
purchased from R & M Marketing, Essex, U.K. Fig. 1 shows the
chemical structure and HPLC chromatograph of amoxicillin, ampicillin
sodium and cloxacillin sodium.
2.2. Antibiotic aqueous solution
Antibiotic aqueous solution was prepared by dissolving specic
amounts of AMX, AMP and CLX in distilled water (dissolved oxygen
8.4 mg/L at 22 C). The aqueous solution characteristics were AMX,
AMP and CLX concentrations 104, 105 and 103 mg/L, respectively,
pH5, COD 520 mg/L, BOD
5
/COD ratio0 and DOC 145 mg/L. The
antibiotic aqueous solution was prepared weekly and stored at 4 C.
2.3. Analytical methods
Antibiotic concentration was determined by a High Performance
Liquid Chromatograph (HPLC) (Agilent 1100 Series) equipped with a
micro-vacuum degasser (Agilent 1100 Series), quaternary pump,
diode array and multiple wavelength detector (DAD) (Agilent 1100
Series) at wavelength 204 nm. The data was recorded by a chemista-
tion software. The column was ZORBAX SB-C18 (4.6 mm150 mm,
5 m) and its temperature was 60 C. The mobile phase was 60%
0.025 M KH
2
PO
4
buffer solution in ultra pure water and 40%
acentonitrile at a ow rate of 0.50 mL/min. Chemical oxygen demand
(COD) was measured according to the Standard Methods [44]. When
hydrogen peroxide was present in the sample, pH was increased to
above 10 to decompose hydrogen peroxide and minimize interference
in COD measurement [45]. A pH meter (HACH Sension 4) and a pH
electrode (HACH platinum series pH electrode model 51910, HACH
Company, USA) was used for pH measurement. Biodegradability was
measured by 5-day biochemical oxygen demand (BOD
5
) test accord-
ing to the Standard Methods [44]. Dissolved oxygen (DO) was
measured by a YSI 5000 dissolved oxygen meter. Bacterial seed was
obtained from a municipal wastewater treatment plant. TOC analyzer
Fig. 1. Chemical structure and HPLC chromatograph of (A) amoxicillin, (B) ampicillin
sodium and (C) cloxacillin sodium.
47 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652
(Model 1010, O & I analytical) was used for determining dissolved
organic carbon (DOC). Ammonia (NH
3
) concentration was measured
by Nessler Method (Method 8038) [46], and nitrate (NO
3

) and
sulphate (SO
4
2
) by Hach Powder Pillow [46].
2.4. Experimental procedure
A 500 mL aliquot of the antibiotic aqueous solution was placed in a
600 mL reactor with the required amount of TiO
2
and was stirred
magnetically at room temperature (222 C). pH of the mixture was
adjusted to the required value by 1 N H
2
SO
4
or 1 N NaOH and the
mixture was kept in the dark for 30 min for dark adsorption before
subjecting to UV irradiation. The source of UV irradiation was a UV
lamp (Spectroline Model EA-160/FE; 230V, 0.17A, Spectronics
Corporation, New York, USA) with a nominal power of 6 W, emitting
radiation at 365 nmand it was placed above the reactor. Samples were
taken at pre-selected time intervals using a syringe and ltered
through a 0.45 m PTFE syringe lter for COD, BOD
5
and DOC
measurement, and through a 0.20 m PTFE syringe lter for determi-
nation of antibiotic concentration by HPLC.
3. Results and discussion
3.1. Effect of UV irradiation
To observe the effect of photolysis of antibiotics due to UVA
irradiation, experiment was conducted at initial AMX, AMP and CLX
concentrations 104, 105, 103 mg/L, respectively and pH 5. By 300-min
UVirradiation, degradationwas 2.9, 3.8and4.9%for AMX, AMP andCLX,
respectively (data not shown). Amoxicillin and cloxacillin show
maximum absorbance at 245 and 250 nm, respectively and they can
absorblight below300 nm(Fig. 2). Therefore, nosignicant degradation
was expected due to UV 365 nm irradiation per se and presumably the
degradation was due to antibiotic hydrolysis. The hydrolysis reaction
would proceed through the attack of the nucleole H
2
O to the -lactam
ring followed by ring opening [39]. Consequently, degradation of the
studied antibiotics whensubjectedto TiO
2
photocatalysis will be mainly
due to the active species produced during the photocatalytic process
(e.g., hydroxyl radical, hole and superoxide ion).
3.2. Effect of TiO
2
concentration
To observe the effect of TiO
2
concentration, initial TiO
2
concen-
tration was varied in the range 0.52.0 g/L. The experimental
conditions were: AMX, AMP and CLX concentrations 104, 105 and
103 mg/L, respectively and pH5. Fig. 3 shows the degradation of AMX,
AMP and CLX. Degradation percent after 300 min irradiation were
42.3, 54.8, 55.8 and 58.7 for AMX (Fig. 3(A)), 33.3, 52.4, 54.3 and 52.4
for AMP (Fig. 3(B)) and 46.6, 58.3, 59.2 and 60.2 for CLX (Fig. 3(C)) at
TiO
2
concentrations 0.5, 1.0, 1.5 and 2.0 g/L, respectively. Degradation
of antibiotics increased with increasing TiO
2
concentration in the
range 0.51.0 g/L. Further increase of TiO
2
concentration above 1 g/L
did not produce signicant improvement in antibiotic degradation.
This may be due to the decreasing light penetration, increasing light
scattering [47], agglomeration, and sedimentation of TiO
2
under high
catalyst concentration [48,49]. The degradation of the antibiotics may
decrease in case of real wastewater due to the presence of interfering
anions (e.g., chloride, bromide, sulphate, phosphate) and/or due to
decrease of light density in turbid wastewater. These results agree
well with previous studies on degradation of bisphenol [50],
chloramphenicol [51], hymatoxylin [33] and paracetamol [8]. Based
on the results, the optimum titanium dioxide concentration for
degradation of amoxicillin, ampicillin and cloxacillin antibiotics in
aqueous solution is 1.0 g/L and it will be used to study the effect of
other parameters. It is worth noting that dissolved oxygen decreased
under the experimental condition from initial value 8.4 to 6.8 mg/L in
300 min. This may be due to the reaction of oxygen with conduction
band electrons to form superoxide ions (
U
O
2

) as in Reaction (5). The


role of dissolved oxygen in photocatalytic degradation is dual. It
accepts a photogenerated electron from the conduction band and
thus promotes the charge separation (minimizing the electron-hole
pair recombination) and it also forms
U
OH radical via superoxide
ion (Reactions (5) and (6)). COD removal percent after 300 min Fig. 2. Absorption spectra of AMX and CLX.
Fig. 3. Effect of TiO
2
concentration on (A) AMX, (B) AMP and (C) CLX degradation.
48 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652
irradiation were 6.2, 9.2, 10.0 and 9.6, respectively. No signicant
improvement in biodegradability (BOD
5
/COD ratio) occurred and the
maximum DOC removal was 2%.
3.3. Effect of pH
To study the effect of initial pH on degradation of AMX, AMP
and CLX, experiments were conducted by varying the pH in the range
311. The experimental conditions were AMX, AMP and CLX con-
centrations 104, 105 and 103 mg/L, respectively, TiO
2
1.0 g/L and
initial COD 520 mg/L. Fig. 4 shows the effect of pH on degradation of
AMX, AMP and CLX. Degradation percent after 300 min irradiation
were 61.2, 54.8, 59.2 and 70.9 for AMX (Fig. 4(A)), 78.1, 52.4, 74.3 and
91.4 for AMP (Fig. 4(B)) and 95.2, 58.3, 81.7 and 100 for CLX(Fig. 4(C))
at pH 3, 5, 8 and 11, respectively.
The effect of pHon antibiotic degradation in terms of COD removal
was also studied. COD removal after 300 min irradiation was 11.7, 9.2,
10.2 and 11.2%, respectively. No signicant improvement in BOD
5
/
COD ratio occurred and the maximum DOC removal was 3%.
The effect of pH on antibiotic degradation can be explained by
taking intoconsiderationthe properties of bothcatalyst andantibiotics
at different pH. For TiO
2
, as pH increases overall surface charge of TiO
2
changes frompositive (pK
a1
=2.6) to negative (pK
a2
=9.0) with point
of zero charge being pH 6.4 [52]. Chemie [53] reported that ionic
amoxicillin species change from positive charge at acidic pH to
negative charge at alkaline pH as shown in Schematic 1. At acidic pH,
both TiO
2
and amoxicillin are positively charged and hence, the
adsorption on the surface of TiO
2
is limited. The high degradation of
antibiotics at acidic pH compared to that at neutral pH may be due to
the hydrolysis of antibiotics as reported by Hou and Poole [54]. At
alkaline pH, both amoxicillin and the TiO
2
are negatively charged and
so repulsive forces between the catalyst and the antibiotics are
developed. High degradation of antibiotics in alkaline condition may
be due to two facts. First is the enhancement of hydroxyl radical
formation at high pH due to the availability of hydroxyl ions on TiO
2
surface that can easily be oxidized to form hydroxyl radical as in
Reaction(3) [8,55,56]. Secondis the hydrolysis of the antibiotics due to
instability of the -lactam ring at high pH [57].
3.4. Kinetics of photocatalytic degradation of amoxicillin, ampicillin and
cloxacillin
To study the kinetics of photocatalytic degradation of amoxicillin,
ampicillin and cloxacillin, experiments were conducted at TiO
2
concentration 1.0 g/L, irradiation time 300 min and pH 11. The con-
centration of AMX, AMP and CLX after 30 min dark adsorption was
taken as the initial AMX, AMP and CLX concentration for kinetic
analysis. Fig. 5 shows the plots of ln ([Antibiotic] / [Antibiotic]
0
) versus
irradiation time for amoxicillin, ampicillin and cloxacillin. The
linearity of the plots as shown in Fig. 5 suggests that the photo-
catalytic degradation approximately followed the pseudo-rst order
kinetics. Degradation of cloxacillin exhibited the highest rate constant
Fig. 4. Effect of pH on (A) AMX, (B) AMP and (C) CLX degradation. Schematic 1. Anionic species of amoxicillin at different pH (Chemie [53]).
49 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652
(0.029 min
1
) followed by amoxicillin (0.007 min
1
) and ampicillin
(0.004 min
1
).
3.5. Effect of H
2
O
2
addition
Addition of H
2
O
2
to TiO
2
suspensions is a well-known procedure
and in many cases leads to an increase in the rate of photocatalytic
oxidation [58,59]. In order to keep the efciency of the added H
2
O
2
, it
is necessary to choose the optimum concentration of H
2
O
2
according
to the type and concentration of the pollutants. H
2
O
2
is considered to
have two functions in the photocatalytic oxidation. It accepts a
photogenerated electron from the conduction band of the semicon-
ductor to form
U
OH radical (Reaction (6)). In addition, it forms
U
OH
radicals according to Reaction (7) [60].
H
2
O
2

U
O
P
2

U
OH OH
P
O
2
7
In order to investigate the effect of H
2
O
2
addition, experiments
were conducted by varying the initial H
2
O
2
concentration in the range
50300 mg/L. The experimental conditions were AMX, AMP and CLX
concentrations 104, 105 and 103 mg/L respectively, TiO
2
1.0 g/L,
ambient pH5 and initial COD 520 mg/L. Ambient pH5 was chosen
because H
2
O
2
decomposes at alkaline pH [45]. Fig. 6 shows the effect
of H
2
O
2
addition on antibiotic degradation in terms of COD removal,
BOD
5
/COD ratio and DOC removal, respectively. COD removal percent
after 300 min illumination was 14.8, 26.3, 23.1, 16.3 and 12.3 at H
2
O
2
concentrations 50, 100, 150, 200 and 300 mg/L, respectively (Fig. 6(A)).
BOD
5
/COD ratio after 300 min illumination was 0.05, 0.10, 0.09, 0.07
and 0.06 at H
2
O
2
concentrations 50, 100, 150, 200 and 300 mg/L,
respectively (Fig. 6(B)). DOC removal percent after 300 min illumina-
tion was 6.4, 13.9, 9.6, 6.9 and 5.3 at H
2
O
2
concentrations 50, 100, 150,
200 and 300 mg/L, respectively (Fig. 6(C)). Maximum COD and DOC
removal percent were achieved at H
2
O
2
concentration 100 mg/L.
Degradation increased as the concentration of H
2
O
2
increased and it
reached the highest value at H
2
O
2
concentration 100 mg/L. Further
increase in H
2
O
2
concentration caused a decrease in COD and DOC
removal and BOD
5
/COD ratio. This may be due to the fact that excess
H
2
O
2
reacts with
U
OHand contributes to the
U
OHand hole scavenging to
formHO
2
U
as inReactions (8) and (9) [61,62]. This agrees well withother
reported studies such as degradation of chloramphenicol [51]. Based on
these results, the optimum H
2
O
2
concentration for antibiotic degrada-
tion is 100 mg/L.
U
OH H
2
O
2
H
2
O HO
U
2
8
H
2
O
2
h

HO
U
2
9
3.6. Degradation of antibiotics and mineralization
To study the effect of H
2
O
2
addition on the degradation of AMX,
AMP and CLX, an experiment was conducted under the following
experimental conditions: TiO
2
1.0 g/L, H
2
O
2
100 mg/L and pH5. Initial
AMX, AMP and CLX concentrations were 104, 105 and 103 mg/L,
respectively and initial COD was 520 mg/L. Fig. 7 shows the effect of
H
2
O
2
addition on AMX, AMP and CLX degradation. Complete
degradation of amoxicillin and cloxacillin were achieved in 20 min,
whereas complete degradation of ampicillin was achieved in 30 min.
TiO
2
photocatalysis would result in the mineralization of organic
carbon, and release of nitrogen and sulphur from the antibiotic
molecule. To assess the degree of mineralization, dissolved organic
carbon (DOC), nitrate (NO
3

), ammonia (NH
3
) and sulphate (SO
4
2
) in
the solution were measured. The nal degradation products were
measured as NO
3

, NH
3
and SO
4
2
concentrations, whereas the small
organic molecules were measured as COD, BOD
5
and DOC. The
experimental conditions were TiO
2
concentration 1.0 g/L, H
2
O
2
concen-
tration 100 mg/L, pH 5, initial AMX, AMP and CLX concentrations 104,
105 and 103 mg/L, respectively and initial COD 520 mg/L. Mineraliza-
tionof organic carbon, andnitrogenandsulphur compounds areveried
Fig. 5. Kinetic analysis of AMX, AMP and CLX degradation.
Fig. 6. Effect of H
2
O
2
addition on (A) COD removal, (B) BOD
5
/COD ratio and (C) DOC
removal, (a) H
2
O
2
concentrations 50, (b) 100, (c) 150, (d) 200 and (e) 300 mg/L.
50 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652
by the results presented in Figs. 8 and 9. Fig. 8 shows the increase
of organic carbon release with irradiation time (DOC removal percent
24.3 and 41.0 at 10 and 24 h, respectively). Fig. 9 shows the evolution
of the ionic species (NO
3

, NH
3
and SO
4
2
) formed during the
degradation of AMX, AMP and AMP. According to the structure of the
antibiotics, each antibiotic has three nitrogen atoms and one sulphur
atom. Photocatalytic transformation of the nitrogen moieties to N
2
,
NH
4
+
, NO
2

, or NO
3

depends onthe initial oxidationstate of nitrogenand


on the structure of the organic molecules [63]. Low et al. [64] reported
that ammonium to nitrate concentration ratio in aliphatic ammines
is higher than that in compounds containing ring nitrogen. For
amoxicillin and ampicillin, each molecule contains three nitrogen
atoms, two of them in aliphatic bond and the other in aromatic bond,
and for cloxacillin, only one atom in aliphatic bond. This indicates that
mineralization of organic nitrogen in case of cloxacillin is more complex
than that of amoxicillin and ampicillin. The results show that the initial
NH
3
concentration was 6.1 mg/L and slightly increased to 7.6 mg/L.
The NO
3

ions gradually increasedfromzero(initial value) to1.2 mg/L in


24 h. Sulphate ions were not detected in the rst 6 h, but were detected
in small amount in the time between 610 h and in high concentration
of 39 mg/L after 24 h. This indicates that the release of sulphur atoms
needs long irradiation time. The results indicate that degradation
of AMX, AMP and CLX antibiotics presumably involves cleavage of
-lactam ringfollowedby subsequent reactions to formcarbondioxide,
water, nitrate, ammonia and sulphate.
4. Conclusions
Degradation of amoxicillin, ampicillin and cloxacillin antibiotics in
aqueous solution by TiO
2
photocatalysis under UVA (365 nm)
irradiation was studied. No signicant degradation occurred by 300-
min UVA irradiation per se. pH had a great effect on antibiotic
degradation and the highest degradation was achieved at pH 11.
Photocatalytic reactions approximately followed a pseudo-rst order
kinetics and the rate constants (k) were 0.007, 0.003 and 0.029 min
1
for amoxicillin, ampicillin and cloxacillin, respectively.
Addition of H
2
O
2
at ambient pH5 and TiO
2
1.0 g/L resulted in
complete degradation of amoxicillin, ampicillin and cloxacillin in
30 min. Dissolved organic carbon (DOC) removal, and nitrate (NO
3

),
ammonia (NH
3
) and sulphate (SO
4
2
) formation during degradation
indicated mineralization of organic carbon, nitrogen and sulphur. UV/
H
2
O
2
/TiO
2
photocatalysis is effective for degradation of amoxicillin,
ampicillin and cloxacillin in aqueous solution.
Acknowledgement
The authors are thankful to the management and authorities of the
Universiti Teknologi PETRONAS for providing facilities for this
research.
References
[1] D.W. Kolpin, E.T. Furlong, M.T. Meyer, E.M. Thurman, Z.S.D. Augg, L.B. Barber, H.T.
Buxton, Environ. Sci. Technol. 36 (2002) 12021211.
[2] P.D. Anderson, V.J. D'Aco, P. Shanahan, S.C. Chapra, M.E. Buzby, V.L. Cunningham, B.M.
Duplessie, E.P. Hayes, F.J. Mastrocco, N.J. Parke, J.C. Rader, J.H. Samuelian, B.W.
Schwab, Environ. Sci. Technol. 38 (2004) 839849.
[3] M. Rabiet, A. Togola, F. Brissaud, J.L. Seidel, H. Budzinski, F. Elbaz-Poulichet,
Environ. Sci. Technol. 40 (2006) 52825288.
[4] M. Carballa, F. Omil, J.M. Lema, M. Llompart, C. Garcia-Jares, I. Rodriguez, M.
Gomez, T. Ternes, Wat. Res. 38 (2004) 29182926.
[5] A. Nikolaou, S. Meric, D. Fatta, Anal. Bioanal. Chem. 387 (2007) 12251234.
[6] P.E. Stackelberg, E.T. Furlong, M.T. Meyer, S.D. Zaugg, A.K. Henderson, D.B.
Reissman, Sci. Total Environ. 329 (2004) 99113.
[7] K. Ikehata, N.J. Naghashkar, M.G. El-Din, Ozone Sci. Eng. 28 (2006) 353414.
[8] L. Yang, L.E. Yua, M.B. Rayb, Wat. Res. 42 (2008) 34803488.
[9] M.V. Walter, J.W. Vennes, Appl. Environ. Microbiol. 50 (1985) 930933.
[10] R. Alexy, T. Kmpel, K. Kummerer, Chemosphere 57 (2004) 505512.
[11] P.A. Pekakis, N.P. Xekoukoulotakis, D. Mantzavinos, Wat. Res. 40 (2006)
1271286.
[12] M.R. Hoffmann, S.T. Martin, W.Y. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995)
6996.
[13] A. Mills, S.L. Hunte, J. Photochem. Photobiol. A Chem. 108 (1997) 135.
[14] D.S. Bhatkhande, V.G. Pangarkar, A.A.C.M. Beenackers, J. Chem, Technol. Biotechnol.
77 (2002) 102116.
[15] K. Kabra, R. Chaudhary, R.L. Sawhney, Ind. Eng. Chem. Res. 43 (2004) 76837696.
[16] I.K. Konstantinou, T.A. Albanis, Appl. Catal. B: Environ. 49 (2004) 114.
[17] W.A. Sadik, A.W. Nashed, A.M. El-Demerdash, J. Photochem. Photobiol. A: Chem.
189 (2007) 135140.
[18] H.K. Singh, M. Saquib, M.M. Haque, M. Muneer, D.W. Bahnemann, J. Mol. Catal.,
A Chem. 264 (2007) 6672.
[19] H. Harada, H. Tanaka, Ultrasonic 44 (2006) 385388.
[20] W. Chu, W.K. Choy, T.Y. So, J. Hazard. Mater. 141 (2007) 8691.
[21] D. Mas, P. Pichat, C. Guillard, F. Luck, Ozone Sci. Eng. 27 (2005) 311316.
[22] C. Guillard, J. Disdier, J.M. Herrmann, C. Lehaut, T. Chopin, S. Malato, J. Blanco, Catal.
Today 54 (1999) 217228.
[23] B. Bayarri, J. Gimnez, D. CurcI, S. Esplugas, Catal. Today 101 (2005) 227236.
[24] Y. Ku, Y.C. Lee, W.U. Wang, J. Hazard. Mater. B138 (2006) 350356.
[25] M. Mrowetz, C. Pirola, E. Selli, Ultrason. Sonochem. 10 (2003) 247254.
[26] M. Qamar, M. Saquib, M. Muneer, Desalination 171 (2004) 185193.
Fig. 7. Effect of H
2
O
2
addition on AMX, AMP and CLX degradation.
Fig. 8. Effect of irradiation time on DOC concentration and removal.
Fig. 9. Effect of irradiation time on NH
3
, NO
3

and SO
4
2
formation.
51 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652
[27] Z. Zainal, L.K. Hui, M.Z. Hussein, Y.H. Tauq-Yap, A.H. Abdullah, I. Ramli, J. Hazard.
Mater. B 125 (2005) 113120.
[28] T. Essam, M.A. Amin, O. El Tayeb, B. Mattiasson, B. Guieysse, Chemosphere 66
(2007) 22012209.
[29] A. Danion, J. Disdier, C. Guillard, O. Pa ss i, N. Jaffrezic-Renault, Appl. Catal. B:
Environ. 62 (2006) 274281.
[30] M.M. Haque, M. Muneer, J. Environ. Mgmt. 69 (2003) 169176.
[31] A.V. Vorontsov, E.N. Savinov, P.G. Smirniotis, Chem. Eng. Sci. 55 (2000) 50895098.
[32] E. Pelizzetti, C. Minero, Electrochim. Acta 38 (1993) 4754.
[33] M. Sioi, A. Bolosis, E. Kostopoulou, I. Poulios, J. Photochem. Photobiol., A Chem. 184
(2006) 1825.
[34] M.N. Abelln, B. Bayarri, J. Gimnez, J. Costa, Appl. Catal., B Environ. 74 (2007)
233241.
[35] M. Audra, J. Andrew, Final Report Submitted to Texas Water Resources Institute,
2003.
[36] A.J. Watkinson, E.J. Murbyc, S.D. Costanzo, Wat. Res. 41 (2007) 41644176.
[37] G. Zhang, S. Ji, B. Xi, Desalination 196 (2006) 3242.
[38] I. Arslan-Alaton, S. Dogruel, J. Hazard. Mater. B112 (2004) 105113.
[39] R. Andreozzi, M. Canterino, R. Marotta, N. Paxeus, J. Hazard. Mater. 122 (2005)
243250.
[40] E. Elmolla, M. Chaudhuri, World Appl. Sci. J. 5 (2009) 5358.
[41] E. Elmolla, M. Chaudhuri, J. Hazard. Mater. 170 (2009) 666672.
[42] E.S. Elmolla, M. Chaudhuri, J. Hazard. Mater. 172 (2009) 14761481.
[43] E.S. Elmolla, M. Chaudhuri. J. Hazard. Mater. in press (2008), doi:10.1016/j.
jhazmat.2009.08.104.
[44] APHA, AWWA, WPCF, Standard Methods for the Examination of Water and
Wastewater, 18th ed., American Public Health Association, American Water Works
Association, Water Pollution Control Federation, Washington, DC, USA, 1992.
[45] I. Talinli, G.K. Anderson, Wat. Res. 26 (1992) 107110.
[46] HACH Handbook, HACH Company, Loveland, CO, USA, 2003.
[47] S.K. Kansal, M. Singh, D. Sud, J. Hazard. Mater. 141 (2007) 581590.
[48] N. San, M. Kilic, Z. Tuiebakhova, Z. Cinar, J. Adv. Oxid. Technol. 10 (2007) 4350.
[49] C.M. So, M.Y. Cheng, J.C. Yu, P.K. Wong, Chemosphere 46 (2002) 905912.
[50] S. Kaneco, M.A. Rahman, T. Suzuki, H. Katsumata, K. Ohta, J. Photochem. Photobiol.,
A Chem. 163 (2004) 419424.
[51] A. Chatzitakis, C. Berberidou, I. Paspalsis, G. Kyriakou, T. Sklaviadis, I. Poulios, Wat.
Res. 42 (2008) 386394.
[52] A.J. Feitz, T.D. Waite, G.J. Jones, B.H. Boyden, P.T. Orr, Environ. Sci. Technol. 33
(1999) 243249.
[53] V.F. Chemie. PhD thesis, Duisburg University, Germany (2005).
[54] J.P. Hou, J.W. Poole, J. Pharm. Sci. 60 (1971) 503.
[55] S.R. Zheng, Q.G. Huang, J. Zhou, B.K. Wang, J. Photochem. Photobiol., A Chem. 108
(1997) 235238.
[56] C. Galindo, P. Jacques, A. Kalt, J. Photochem. Photobiol., A Chem. 130 (2000) 3547.
[57] A.D. Deshpande, K.G. Baheti, C.N.R. Chatterjee, Current Sci. 87 (2004) 16841695.
[58] S. Malato, J. Blanco, M. Maldonado, P. Fernandez-Ibanez, A. Campos, Appl. Catal. B:
Environ. 28 (2000) 163174.
[59] I. Poulios, E. Micropoulou, R. Panou, E. Kostopoulou, Appl. Catal., B Environ. 41
(2003) 345355.
[60] M. Kositzi, A. Antoniadis, I. Poulios, I. Kiridis, S. Malato, Sol. Ener. 77 (2004)
591600.
[61] H. Zhao, S. Xu, J. Zhong, X. Bao, Catal. Today 9395 (2004) 857861.
[62] M.A. Behnajady, N. Modirshahla, R. Hamzavi, J. Hazard. Mater. B133 (2006)
226232.
[63] P. Calza, E. Pelizzetti, C. Minero, J. Appli. Electrochem. 35 (2005) 665673.
[64] G.K.C. Low, S.R. McEvoy, R.W. Matthews, Environ. Sci. Technol. 25 (1991) 460467.
52 E.S. Elmolla, M. Chaudhuri / Desalination 252 (2010) 4652

Você também pode gostar