Sei sulla pagina 1di 13

This article was downloaded by: [University of York]

On: 19 August 2014, At: 03:53


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Toxicological & Environmental


Chemistry
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gtec20

Sulfamethoxazole removal from


polluted water by immobilized
photocatalysis
ab

A. Alatrache , A. Cortyl , P. Arnoux , M.N. Pons & O. Zahraa

Laboratoire Sciences et Gnie des Procds (LRGP) CNRS, UMR


7274, ENSIC, Universit de Lorraine, Nancy Cedex, France
b

Laboratoire RMN Physique 01 UR 13-04, Ecole Suprieure des


Sciences et Techniques de Tunis, Tunis, Tunisia
Published online: 14 Aug 2014.

To cite this article: A. Alatrache, A. Cortyl, P. Arnoux, M.N. Pons & O. Zahraa (2014):
Sulfamethoxazole removal from polluted water by immobilized photocatalysis, Toxicological &
Environmental Chemistry, DOI: 10.1080/02772248.2014.942308
To link to this article: http://dx.doi.org/10.1080/02772248.2014.942308

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the
Content) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &

Downloaded by [University of York] at 03:53 19 August 2014

Conditions of access and use can be found at http://www.tandfonline.com/page/termsand-conditions

Toxicological & Environmental Chemistry, 2014


http://dx.doi.org/10.1080/02772248.2014.942308

Sulfamethoxazole removal from polluted water by immobilized


photocatalysis
A. Alatrachea,b, A. Cortyla, P. Arnouxa, M.N. Ponsa and O. Zahraaa*
a
Laboratoire Sciences et G
enie des Proc
ed
es (LRGP) CNRS, UMR 7274, ENSIC, Universit
e de
erieure des
Lorraine, Nancy Cedex, France; bLaboratoire RMN Physique 01 UR 13-04, Ecole Sup
Sciences et Techniques de Tunis, Tunis, Tunisia

Downloaded by [University of York] at 03:53 19 August 2014

(Received 13 February 2014; accepted 17 June 2014)


The photocatalytic activity of TiO2 deposits (Degussa P25 and Millennium PC500)
has been studied using sulfamethoxazole (SMX) as a model water pollutant and a UV
fluorescent lamp as a light source (365 nm). Both catalysts have shown very similar
properties in the photocatalytic degradation of SMX. Special attention has been given
to the effect of the irradiation time, pH, and pollutant concentration. No mass-transfer
limitations are observed. The degradation of SMX is accelerated at low concentration,
and the photocatalytic degradation kinetics obey the LangmuirHinshelwood model,
allowing the adsorption and apparent rate constants to be determined for both
catalysts.
Keywords: advanced oxidation processes (AOPs); photocatalytic degradation; LangmuirHinshelwood; antibiotics

Introduction
Discharges of industrial and domestic effluents (even when treated for C/N/P macropollution) into surface waters (rivers and lakes) introduce various micropollutants to the water
cycle. The intensification of agriculture (including fish farming) also introduces pollutants, and groundwater can be affected by infiltration through soil. Many micropollutants
are found ranging from heavy metals to organic compounds (Halling-Srensen et al.
1998). Despite efforts to provide high-quality drinking water using water pollution control, the facilities currently available are unable to remove all traces of pollutants in water
resources (Suarez et al. 2008; Koester et al. 2012). Antibiotics are among these pollutants;
they enter the aquatic environment from different sources such as wastewater discharged
from production facilities and urban treatment systems and runoff from farms. Source
points in cities are hospitals and other care units, but diffuse sources should not be overlooked (home use) (K
ummerer 2009). Although some countries limit the use of antibiotics as growth factors in farming, therapeutic use is still widespread (Trevisi et al. 2014).
Besides the direct lethal effect, these chemicals can have bacteria involved in biological
wastewater treatment (activated sludge, biofilms) (Louvet et al. 2010), they can also
induce antibiotic resistance (Li et al. 2010; Rizzo et al. 2013). Antibiotic (and biocide)
resistance genes can be transferred to pathogenic and non-pathogenic bacteria in sewage
transport and treatment systems as well as aquatic ecosystems. Among these antimicrobials, sulfonamides are extensively used in both human and veterinary medicine. Sulfamethoxazole (SMX) is the most commonly employed antibiotic in the latter field: this
*Corresponding author. Email: orfan.zahraa@univ-lorraine.fr
2014 Taylor & Francis

Downloaded by [University of York] at 03:53 19 August 2014

A. Alatrache et al.

Figure 1. Principle of the photocatalytic process.

powerful antibiotic is often found in the aquatic environment (Baran, Sochacka, and
Wardas 2006; Haller et al. 2002).
In order to eliminate micropollution, water treatment can rely on the adsorption of
pollutants on the surface of an appropriate adsorbent, and the mineralization of these pollutants by oxidation. Adsorption on activated carbon is occasionally insufficient. In addition, the adsorption requires the treatment of the used adsorbent to eliminate the adsorbed
pollutant. Oxidative treatment by chlorine or chlorine dioxide leads to the formation of
non-friendly, harmful by-products (Della-Greca et al. 2009). Advanced oxidation processes (AOPs) that use hydrogen peroxide or ozone are more efficient than other processes, but may be costly (Coleman, Chiang, and Amal 2005; Ismal, Bousselmi, and
Zahraa 2011). An alternative emerging cheaper AOP technology is heterogeneous photocatalysis (Ollis, Pelizzetti, and Serpone 1991), where most organics compounds are ultimately transformed into non-toxic inorganic species under the action of light and oxygen.
The advantages of this process include the absence of soluble additives and the possibility
of using solar light. The photocatalyst is mainly a semiconductor, oxide, or sulfide. Titanium dioxide (TiO2) is a promising photocatalyst for the elimination of toxic and bioresistant organic and inorganic compounds from wastewater by the transformation of these
compounds into harmless species. Absorption of a photon in the near-UV range promotes

C
an electron, ecb
, to the conduction band, which produces a hole, hvb
, in the valence band.
If separated, these two species can migrate to the catalyst surface and act as a reducer
(oxygen reduction of a superoxide ion) and an oxidant (water oxidation of a hydroxyl radical, for example, or direct oxidation of the reactant), thus regenerating the catalyst electronic population. Highly reactive species such as hydroxyl radicals can react on
adsorbed organic molecules and abstract a hydrogen atom, and therefore, induce oxidation, as shown in Figure 1.
In the present work, the feasibility of SMX degradation by photocatalysis was evaluated using two types of immobilized photocatalysts, and the influence of several parameters affecting the performance of photocatalysts was determined. Finally, a kinetic
expression for the heterogeneous photocatalytic degradation in laboratory reactor was
developed.

Downloaded by [University of York] at 03:53 19 August 2014

Toxicological & Environmental Chemistry

Figure 2. Experimental setup: (1) cover, (2) fluorescent ultraviolet lamp, (3) reactor, (4) plate with
immobilized photocatalyst, (5) PTFE tubing, (6) peristaltic pump, (7) reservoir, and (8) magnetic
stirrer.

Materials and methods


SMX (99% purity) was purchased from Sigma Chemical Co. (Saint Quentin Fallavier,
France), and used without further purification. The maximum absorbance wavelength of
this chemical was identified using a Perkin ElmerUV/Vis Lambda EZ210 spectrophotometer at 262 nm. For both types of photocatalysts (titanium dioxide Degussa P25 and
Millennium PC500, which were given by the respective manufacturers), the preparation
protocol has been presented in detail elsewhere (Ould-Mame, Zahraa, and Bouchy 2000).
The acidic suspension of TiO2 in ethanol was spread on a glass plate before drying and firing at 470  C. For each catalyst, three layers of TiO2 were deposited on the plate to obtain
the best performance in terms of total photon absorption. Photocatalytic measurements
were carried out on a laboratory reactor designed for testing 40 500 mm2 flat glass
plates with the deposited catalyst film (Zahraa et al. 1999). The total volume of the aqueous solution was 0.25 L. The solution was continuously circulated with a peristaltic
pump, flowing freely over the catalytic surface and stirred in the 0.2-L reservoir. A schematic diagram of the experimental setup is shown in Figure 2.
The system could be considered to be constantly air-equilibrated. The circulating flow
rate could be varied up to 400 mL min1. This flow rate was checked in preliminary
experiments, and almost no effect was observed when the flow rate was between 100 and
400 mL min1. As the extent of adsorption or photocatalytic degradation was very small
during a single pass (0.5 min), the whole system could be considered a perfectly stirred
closed reactor with respect to the solution. Irradiation of the photocatalyst was carried out
using a MAZDA-TFN18 18 W UV fluorescent lamp emitting approximately 365 nm (95%
of the light emitted below 400 nm) or a visible light lamp (Osram L18W/640 with 90% of
the light emitted above 400 nm). Both of these lamps were positioned parallel to the plate
at a distance of 50 mm. The irradiation intensity versus the distance to the UV lamp was
measured with a Vilber Lourmat VLX365 radiometer and was equal to 1.45 mW cm2 at
a distance of 50 mm (5% depending upon the location). The SMX concentration was
monitored during the reaction in the aqueous phase by UV/Vis spectrophotometry and by
an high performance liquid chromatography (HPLC) equipped with a Shimadzu SPD20AT spectrophotometer detector adjusted to the wavelength of 262 nm corresponding to

Downloaded by [University of York] at 03:53 19 August 2014

A. Alatrache et al.

Figure 3. Illustration of the photocatalytic process ([SMX]0 D 25 mg L1, flow rate D 220 mL
min1, catalyst: PC500).

the maximal absorption of SMX. A Grace Smart RP 18 5 mm, 250 2.1 mm column was
used, and a mixture of methanol (50%) and water (50%) was used for elution.

Results and discussion


In this study, each experiment was repeated three times, and the experimental values in
the figures represent the mean values of three replicates with standard deviation of <5%.
Photocatalytic process
A preliminary experiment shown in Figure 3 demonstrates the photocatalytic nature of the
degradation process. In this experiment, three runs were carried out: (1) with UV only, (2)
with TiO2 and visible light, and (3) with TiO2 and UV light. In all of the runs and prior to
t D 0, the reactive mixture circulated in the photocatalytic reactor for a period of 60 min
in the dark to ensure perfect mixing and adsorption in the presence of the photocatalyst.
The absence of significant adsorption at this step should be mentioned. The SMX concentration was then at the initial value, [SMX]0. The relative decrease in the concentration
([SMX]/[SMX]0) was then monitored under the three conditions described above.
Accordingly, the photocatalytic process occurs at room temperature in the simultaneous
presence of catalyst and photons. These conditions could lead to the total mineralization
of the antibiotic. Several researchers have focused their work on this subject in batch
systems, where suspensions of TiO2 nanoparticles were irradiated (Hu et al. 2007; Javier
Benitez et al. 2011). Other researchers, as in the present study, have immobilized the catalyst to avoid the post-treatment separation step in the case of an industrial application
(Coleman et al. 2000). However, the immobilization of the nanoparticles can limit the
surface contact. Consequently, lower activity is expected, which could be overcome with
modifying the catalytic material (Coleman, Chiang, and Amal 2005; Ismal, Bousselmi,
and Zahraa 2011). The photocatalytic degradation of SMX was then monitored at
different irradiation times, as shown in Figure 4. The decrease in the SMX concentration
and the absence of the eventual formation of by-products are observed, except that the

Downloaded by [University of York] at 03:53 19 August 2014

Toxicological & Environmental Chemistry

Figure 4. Variation of SMX solution absorbance spectra as a function of the irradiation time
for P25.

slight absorbance at 225 nm, which disappeared at longer time of irradiation. However,
variations in the absorbance in the UV region can indicate the possible production and
disappearance of by-products.
Abellan et al. (2007) identified five degradation products by liquid chromatographymass spectrometry (LC-MS) that can be formed during the photocatalytic degradation
of SMX. Nasuhoglu, Yargeau, and Berk (2011) have been working on some of these
by-products, such as sulfanilic acid and 3-amino-5-methylisoxazole with an absorbance
at 225 nm. These authors have observed good removal of these by-products as indicated
by measurements of total organic carbon.

The effect of the catalyst type


To compare the catalytic activities of P25 and PC500, experiments were conducted for
each catalyst with an identical initial concentration of SMX (25 mg L1). Figure 5(A)
shows that P25 was significantly more active than PC500. After 420 min, the photocatalytic degradation was 94% over P25 and 79% over PC500. This observation was verified
when the absorption measurements were performed using a spectrophotometer equipped
with integrating sphere. The absorbance spectra of both types are given in Figure 5(B).
The TiO2 P25 catalyst exhibits minimal absorption in the visible region; however, under
visible irradiation, no photocatalytic activity was detected.
The difference in the activity observed above can be explained by the presence of
rutile in P25 (the material is 70% anatase and 30% rutile), where PC500 is composed of
>99% anatase. This result is in accordance with the observations of Bickley et al. (1991),
who ascribed the higher activity of P25 to its structure. Moreover, Martin et al. (1994)
attributed the superiority of P25 to the slower electron/hole recombination that occurs on
the catalyst surface. Despite the efficiency of the P25 catalyst, the following experiments
were conducted with the two types of catalysts to see if improvements could be obtained
for PC500 under certain operating conditions.

Downloaded by [University of York] at 03:53 19 August 2014

A. Alatrache et al.

Figure 5. (A) Influence of the catalyst type on the variation of [SMX]/[SMX]0 versus time.
[SMX]0 D 25 mg L1, flow rate D 220 mL min1. (B) Absorbance spectra of three layers of P25
and PC500 deposited on glass plates.

The effect of pH
To study the influence of pH, experiments were conducted at different initial pH values
(5, 7, and 9). The initial pH was adjusted with NaOH and HNO3, the initial SMX concentration of the solution was 25 mg L1, and the flow rate was set at 220 mL min1. As
expected, this investigation shows a significant decrease in the initial degradation rate
(r0) for both catalysts when the pH increases from 5 to 7: r0 (pH D 7) D 17% r0 (pH D 5)
for P25, and r0 (pH D 7) D 50% r0 (pH D 5) for PC500. The literature indicates that the
SMX molecule contains one basic amine group (NH2) and one acidic amide group
(NH), and therefore, this molecule has two pKa values (1.7 and 5.6) (Lucida, Parkin,
and Sunderland 2000). Xekoukoulotakis et al. (2011) showed the variation in the SMX
charge with different pH values. Furthermore, at pH values greater than approximately
6.7, the titania surface becomes negatively charged (Fernandez-Iba~nez et al. 2003). At
near-neutral/slightly alkaline pH, both the SMX molecule and the catalyst surface are
negatively charged, and the repulsion between them is enhanced. Therefore, a reduction
in the photoadsorption of SMX onto the catalyst surface is expected. Additionally,
Abellan et al. (2007) concluded that the SMX degradation remains constant and that the
charge of neither the titania surface nor the SMX molecule influences the activity of an
aqueous suspension of TiO2. At the end of each experiment, the pH was measured. An
almost systematic decrease in pH is observed compared to the initial solution pH (especially for pH 7 and pH 9). This decrease in pH could result from the formation of acidic
species throughout the reaction due to mineralization by SMX (Abellan et al. 2007).

Downloaded by [University of York] at 03:53 19 August 2014

Toxicological & Environmental Chemistry

Figure 6. Influence of the flow rate on the initial rate of photocatalytic degradation of SMX (r0) for
P25 deposit.

The effect of the flow rate


In heterogeneous catalysis, investigations of the kinetics of reactions are focused on two
considerations: physical kinetics that are determined by the mass-transfer step, and chemical kinetics that are determined by the reaction and chemisorption steps. Generally,
chemical kinetics determine the observed rate because mass transfer is a fast process and
diffusion limitations are avoided.
Experiments were conducted to determine the existence of external mass-transfer limitations; the reaction rate was then plotted against the flow rate. Figure 7 shows the variation in the SMX degradation rate with the TiO2 P25 catalyst.
As shown in Figure 6, the flow rate has a slight effect on the rate r0 of photocatalytic
degradation, which means that an external mass-transfer limitation does not intervene.
However, under the standard conditions of the present study, at the maximum flow rate
the rate r0 is nearly independent of the flow rate, indicating that the external mass transfer
is fast and has practically no effect on the reaction kinetics.
In addition, in our experimental setup, the mass-transfer coefficient can be assumed to
be proportional to the flow rate4/5 (Sissom and Pitts 1972). Thus, increasing the flow rate
from 110 to 440 mL1min would increase the rate by a factor of approximately 3, if the
rate was dominated by external mass transfer. However, the experimental increase in the
rate is only 20%, which corroborates the proposed effect of the external mass transfer on
the reaction kinetics.

The effect of the SMX concentration


The effect of varying the initial SMX concentration on the performance of both the photocatalysts was studied in the range of 1.5100 mg L1 at a flow rate of 220 mL min1. The
results are shown in Figures 7(A) and 7(B). As illustrated in this figure, the rate of degradation of SMX decreases when the initial SMX concentration is increased from 1.5 to
100 mg L1. The degradation efficiency is very low at the highest concentrations for the
PC500 catalyst. This behavior can be expected in our system because the concentrated
reaction solution mimics a filter that limits the photons reaching the catalyst surface. Similar results were observed when Abellan et al. (2007) and Xekoukoulotakis et al. (2011)
monitored the degradation of SMX in the presence of TiO2 particles. In addition, the

Downloaded by [University of York] at 03:53 19 August 2014

A. Alatrache et al.

Figure 7. (A) Effect of SMX concentration on the activity of TiO2 P25. (B) Effect of SMX concentration on the activity of TiO2 PC500.

exponential form of the curves clearly indicates that the kinetics of the photocatalytic
degradation of SMX are a first-order reaction. This result is in agreement with the simple
LangmuirHinshelwood (LH) model. In the LH model, the degradation reaction follows
a rate-determining step where an adsorbed molecule reacts with a reactive transient such
as an OH radical or a hole. If this attack is a minor process in the disappearance of the
reactive transient, the rate of reaction is proportional to the surface coverage by the pollutant. Assuming a simple Langmuir model of adsorption, the rate of reaction is given by the
relationship described by Scacchi et al. (1996):
r kdeg

KLH SMX
1 KLH SMX

(1)

where KLH is the adsorption constant, [SMX] is the SMX concentration in the aqueous
phase, and kdeg is the apparent kinetics constant, which depends on various physicochemical parameters including the irradiation conditions. Other researchers, Aramenda et al.
(2005) and Chong et al. (2010), also considered this model in their interpretation of the
kinetics of photocatalytic degradation. Relationship 1 can be rewritten in a linear form as
1
1
1
1
r0 kdeg KLH:  SMX C kdeg .
0

Downloaded by [University of York] at 03:53 19 August 2014

Toxicological & Environmental Chemistry

Figure 8. The LangmuirHinsheloowd model for SMX degradation on P25 catalyst, C0 in mg.L1
and r0 in mg.L1 min1, inset figure, experimental data, model.

The model parameters were then estimated by plotting 1/r0 versus 1/[SMX]0. The data
are consistent with this linear relationship, and the initial rate is modeled (see inset) as
shown in Figure 8.
The values for KLH and kdeg derived from the LH kinetic model are presented in
Table 1. These values indicate that both catalysts offer SMX the same accessibility
toward their surfaces because the adsorption constants are of the same order of magnitude. However, the apparent degradation rate constant of P25 is threefold higher the rate
constant of PC500, which proves once more the better performance of P25 in comparison
to the performance of PC500.
Conclusions
SMX is efficiently degraded by photocatalysis on immobilized TiO2. The structure of
Degussa P25 causes this catalyst to exhibit higher activity than Millennium PC500. Furthermore, experimental parameters such as irradiation time and pH affect the performance
of the catalysts. In contrary, the flow rate variations of the solution do not modify the degradation of SMX. It is interesting to note that this degradation is accelerated at low concentration. The kinetics agree with the LH model, and the adsorption constant and
apparent rate constant from this model have been determined for both catalysts. This

Table 1. LangmuirHinshelwood model kinetic constants.

kdeg (mg L1 min1)


KLH (L mg1)
R2

P25

PC500

0.36
0.032
0.96

0.13
0.036
0.85

10

A. Alatrache et al.

experimental study could provide useful information for the design of an industrial elimination processing unit.

Downloaded by [University of York] at 03:53 19 August 2014

References
Abellan M.N., B. Bayarri, J. Gimenez, and J. Costa. 2007. Photocatalytic Degradation of Sulfamethoxazole in Aqueous Suspension of TiO2. Applied Catalysis B: Environmental 74:
233241.
Aramenda M.A., A. Marinas, J.M. Marinas, J.M. Moreno, and F.J. Urbano. 2005. Photocatalytic
Degradation of Herbicide Fluroxypyr in Aqueous Suspension of TiO2. Catalysis Today 101:
187193.
Baran W., J. Sochacka, and W. Wardas. 2006. Toxicity and Biodegradability of Sulfonamides and
Products of Their Photocatalytic Degradation in Aqueous Solutions. Chemosphere 65:
12951299.
Bickley R.I., T. Gonzalez-Carreno, J.S. Lees, L. Palmisano, and R.J.D. Tilley. 1991. A Structural
Investigation of Titanium-Dioxide Photocatalysts. Journal of Solid State Chemistry 92:
178190.
Chong M.N., B. Jin, C.W.K. Cho, and C. Saint. 2010. Recent Developments in Photocatalytic
Water Treatment Technology: A Review. Water Research 44: 29973027.
Coleman H.M., K. Chiang, and R. Amal 2005. Effect of Ag and Pt on Photocatalytic Degradation
of Endocrine Disrupting Chemicals in Water. Chemical Engineering Journal 113: 6572.
Coleman H.M., B.R. Eggins, J.A. Byrne, F.L. Palmer, and E. King. 2000. Photocatalytic Degradation of 17b-oestradiol on Immobilized TiO2. Applied Catalysis B: Environmental 24: 15.
Della-Greca M., M.R. Iesce, P. Pistillo, L. Previtera, and F. Temussi. 2009. Unusual Products of
the Aqueous Chlorination of Atenolol. Chemosphere 74: 730734.
Fernandez-Iba~nez P., J. Blanco, S. Malato, and F.J. de las Nieves. 2003. Application of the Colloidal Stability of TiO2 Particles for Recovery and Reuse in Solar Photocatalysis. Water Research
37: 31803188.
Haller M.Y., S.R. Muller, C.S. McArdell, A.C. Alder, and M.J.F. Suter. 2002. Quantification of
Veterinary Antibiotics (Sulfonamides and Trimethoprim) in Animal Manure by Liquid ChromatographyMass Spectrometry. Journal of Chromatography A 952: 111120.
Halling-Srensen B., S. Nors Nielsen, P.F. Lanzky, F. Ingerslev, H.C. Holten L
utzhoft, and S.E.
Jrgensen. 1998. Occurrence, Fate and Effects of Pharmaceutical Substances in the Environment. Chemosphere 36: 357393.
Hu L., P.M. Flanders, P.L. Miller, and T.J. Strathmann. 2007. Oxidation of Sulfamethoxazole and
Related Antimicrobial Agents by TiO2 Photocatalysis. Water Research 41: 26122626.
Ismal M., L. Bousselmi, and O. Zahraa. 2011. Photocatalytic Behavior of WO3-Loaded TiO2 Systems in the Oxidation of Salicylic Acid. Journal of Photochemistry and Photobiology A:
Chemistry 222: 314322.
Javier Benitez F., J.L. Acero, F.J. Real, G. Roldan, and F. Casas. 2011. Comparison of Different
Chemical Oxidation Treatments for the Removal of Selected Pharmaceuticals in Water
Matrices. Chemical Engineering Journal 168: 11491156.
Koester S., S. Beier, F.F. Zhao, Q. Sui, G. Yu, and J. Pinnekamp. 2012. Organic Trace Pollutants in
the Aquatic Environment-Regulatory and Technical Problem-Solving Approaches in Germany
and China. Water Science and Technology 66: 942952.
Kummerer K., 2009. Antibiotics in the Aquatic Environment  a Review  Part II. Chemosphere
75: 435441.
Li, D., T. Yu, Y. Zhang, M. Yang, Z. Li, M. Liu, and R. Qi. 2010. Antibiotic Resistance
Characteristics of Environmental Bacteria from an Oxytetracycline Production Wastewater
Treatment Plant and the Receiving River. Applied and Environmental Microbiology 76:
34443451.
Louvet J.N., Y. Heluin, R. Attik, D. Dumas, O. Potier, and M.N. Pons. 2010. Erythromycin Toxicity on Activated Sludge: Assessment via Batch Experiments and Microscopic Techniques (Epifluorescence and CLSM). Process Biochemistry 45 (11): 17871794.
Lucida H., J.E. Parkin, and V.B. Sunderland. 2000. Kinetic Study of the Reaction of Sulfamethoxazole and Glucose Under Acidic Conditions - I. Effect of pH and Temperature. The International Journal of Pharmaceutics 202: 4761.

Downloaded by [University of York] at 03:53 19 August 2014

Toxicological & Environmental Chemistry

11

Martin S.T., H. Herrmann, W. Choi, and M.R. Hoffmann. 1994. Time-Resolved Microwave Conductivity. I. TiO2 Photoreactivity and Size Quantization. The Journal of the Chemical Society,
Faraday Transactions 90 (21): 33153322.
Nasuhoglu D., V. Yargeau, and D. Berk. 2011. Photo-Removal of Sulfamethoxazole (SMX) by
Photolytic and Photocatalytic Processes in a Batch Reactor Under UV-C Radiation (max D
254 nm). Journal of Hazardous Materials 186: 6775.
Ollis D.F., E. Pelizzetti, and N. Serpone. 1991. Photocatalyzed Destruction of Water Contaminants. Environmental Science & Technology 25: 15231529.
Ould-Mame S.M., O. Zahraa, and M. Bouchy. 2000. Photocatalytic Degradation of Salicylic Acid
on Fixed TiO2  Kinetic Studies. Journal of Advanced Oxidation Technologies 2: 5966.
Rizzo L, C. Manaia, C. Merlin, T. Schwartz, C. Dagot, M.C. Ploy, I. Michael, and D. Fatta-Kassinos. 2013. Urban Wastewater Treatment Plants as Hotspots for Antibiotic Resistant Bacteria
and Genes Spread into the Environment. A Review. Science of the Total Environment 447:
345360.
Scacchi G., M. Bouchy, J.F. Foucaut, and O. Zahraa. 1996. Cin
etique et catalyse, Collection G
enie
des proc
ed
es de l
ecole de Nancy [Kinetic and Catalysis]. Paris: Ed. TEC & DOCLavoisier.
Sissom L.E., and D.R. Pitts. 1972. Elements of Transport Phenomena. New York: McGraw-Hill.
Suarez, S., M. Carballa, F. Omil, and J.M. Lema. 2008. How Are Pharmaceutical and Personal
Care Products (PPCPs) Removed from Urban Wastewaters? Reviews in Environmental Science and Biotechnology 7: 125138.
Trevisi E., A. Zecconi, S. Cogrossi, E. Razzuoli, P. Grossi, and M. Amadori. 2014. Strategies for
Reduced Antibiotic Usage in Dairy Cattle Farms. Research in Veterinary Science 96:
229233.
Xekoukoulotakis N.P., C. Drosou, C. Brebou, E. Chatzisymeon, E. Hapeshib, D. Fatta-Kassinos,
and D. Mantzavinos. 2011. Kinetics of UV-A/TiO2 Photocatalytic Degradation and Mineralization of the Antibiotic Sulfamethoxazole in Aqueous Matrices. Catalysis Today 161:
163168.
Zahraa O., C. Dorion, S.M. Ould-Mame, and M. Bouchy. 1999. Titanium Dioxide Deposit Films
for Photocatalytic Studies of Water Pollutants. Journal of Advanced Oxidation Technologies
4: 4046.

Potrebbero piacerti anche