Vous êtes sur la page 1sur 11

Journal of Environmental Management 271 (2020) 110964

Contents lists available at ScienceDirect

Journal of Environmental Management


journal homepage: http://www.elsevier.com/locate/jenvman

Research article

Photocatalytic degradation of model pharmaceutical pollutant by novel


magnetic TiO2@ZnFe2O4/Pd nanocomposite with enhanced photocatalytic
activity and stability under solar light irradiation
Najmeh Ahmadpour a, Mohammad Hossein Sayadi a, *, Sara Sobhani b, Mahmood Hajiani a
a
Department of Environmental Engineering, Faculty of Natural Resources and Environment, University of Birjand, Birjand, Iran
b
Department of Chemistry, College of Sciences, University of Birjand, Birjand, Iran

A R T I C L E I N F O A B S T R A C T

Keywords: In the last decades, the use of magnetic nanocomposites as a catalyst was considered for removal of organic
Composite TiO2@ZnFe2O4/Pd photocatalyst pollutants due to its easy separation. Therefore, initially, TiO2@ZnFe2O4/Pd nanocomposite was prepared and
Batch and continuous systems- visible light then used in the photodegradation of diclofenac under direct solar irradiation in the batch and continuous
photocatalysis
systems. The structure, morphology and other specifications of produced nanocatalyst were determined via XRD,
Photocatalytic degradation- diclofenac
VSM, FESEM/EDX, FTIR, GTA, UV–Vis, Zeta potential, XPS and ICP-OES. The effective factors on diclofenac
photocatalytic degradation
Magnetic characteristics of the photocatalyst removal via nanophotocatalyst viz. pH, catalyst concentration, initial concentration of diclofenac, and flow rate
and column length on diclofenac photodegradation were studied. Based on the results, the optimal rate for pH,
catalyst concentration, and initial concentration of diclofenac was 4, 0.03 g/l and 10 mg/l respectively. Pd-
coated TiO2@ZnFe2O4 magnetic photocatalyst had higher photocatalytic activity in diclofenac photo­
degradation in relation to ZnFe2O4 and TiO2@ZnFe2O4 under solar light irradiation. The findings showed that
after five recycles, the photocatalytic efficiency did not show much reduction i.e. the removal efficiency from
86.1% in the first cycle reduced only to 71.38% in the last cycle. Likewise, in this study, with flow rate reduction
and column length increase diclofenac degradation rate increased.

1. Introduction diclofenac (DCF) has been determined at its highest concentration i.e.
28.4 μg/l in the surface waters. Moreover, in the ground waters it has
In the past few years, the growing of industrialization and excessive been reported at highest concentrations of 0.59 μg/l (De Oliveira et al.,
utilization of available resources by humans has emerged water pollu­ 2018). Therefore, numerous techniques were used to remove pharma­
tion as one of the global environmental issues (Hasija et al., 2019). Entry ceutical pollutants from wastewater (Guo et al., 2018). Among them, the
of pharmaceutical pollutants to the environmental resources viz. surface Advanced Oxidation Process (AOP) is an effective technique in which a
waters, ground waters, sediments, soil and even drinking water, has semiconductor catalyst can with use of natural energy i.e. solar irradi­
created a serious danger for the humans and freshwater besides marine ation eliminate the pharmaceutical pollutants (Zuorro et al., 2013).
water ecosystem. The pharmaceutical industries, hospitals and improper Advanced oxidation processes such as photocatalysis (Calza et al.,
disposal of pharmaceuticals are the main resources of pharmaceutical 2006), UV (Kim and Tanaka, 2009), UV/H2O2 (Vogna et al., 2006),
pollutants. Among the pharmaceutical pollutants diclofenac 2-[2-[(2, Photo-Fenton (Ravina et al., 2002; Perez-Estrada et al., 2005), ozonation
6-dichlorophenyl)amino]phenyl]acetic acid is a non-steroidal anti-in­ (Vieno et al., 2007) and sonolysis (Naddeo et al., 2010) for DCF removal
flammatory drug that is used as antipain, anticancer, and anti-rheumatic from polluted water has been reported wherein among these processes,
combinations and is one of the most commonly used drugs that have photocatalysis is an effective technique due to low cost, good stability,
potential toxic effects (Zhao et al., 2017). Considering the nonpolar eco-friendly nature, and energy efficiency wherein a semiconductor
putative nature, the distribution and release of diclofenac in the envi­ catalyst can via use of natural energy of solar irradiation eliminate the
ronment takes place naturally via aquatic migration and its accumula­ organic pollutants (Sharma et al., 2019). Basically, photocatalysis
tion in the food chain (Acu~ na et al., 2015) and in the recent years, mechanism is transfer of photon energy to chemical energy. Here, a

* Corresponding author.
E-mail address: mh_sayadi@birjand.ac.ir (M.H. Sayadi).

https://doi.org/10.1016/j.jenvman.2020.110964
Received 5 January 2020; Received in revised form 5 June 2020; Accepted 12 June 2020
Available online 29 June 2020
0301-4797/© 2020 Elsevier Ltd. All rights reserved.
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

photocatalyst plays an important role that using solar light absorption synthesized using the photodeposition technique. Initially 0.0052 g Pd
(hv), leading to creation of electron-hole recombination pair, separation (NO3)2 2H2O was dissolved in 40 ml DW and later 0.1 g TiO2@ZnFe2O4
of electrons in conduction band and oxidizing holes in valence band in was added to it. To obtain a homogenous solution of magnetic nano­
energy level platform (Hasija et al., 2020). Among the different cata­ structures in the solvent, it was sonicated for 20 min. Later, the mixture
lysts, semiconductor photocatalysts have exhibited a significant per­ was transferred to a Quartz tube; 0.4 ml methanol was added and sub­
formance against the degradation of organic and inorganic pollutants. jected for 10 h under ultraviolet light. In this stage, it can be stated that
TiO2 is one of the commercial and efficient photocatalysts that is used palladium ions are reduced to metallic palladium and is positioned at
for the degradation of most of the pollutants, due to its flexibility, ease of TiO2 level, later Pd are trapped at electron sites and are stored as an
synthesis, ability to control and good stability (Czili and Horv�ath, 2009). oxide.
However, the two inherent weaknesses of TiO2 viz. large bandgap (EV
3.2) and high recombination rate, limits its photoresponse in the UV 2.3. The characterization of TiO2@ZnFe2O4/Pd nanoparticles
region (4–5% of solar spectrum) and quantum efficiency reduces the
photocatalytic reactions, besides its separation after the process is an The crystalline structure of the samples was determined via X-ray
issue (Lim et al., 2017; Chen and Burda, 2008). Several attempts to in­ Diffraction (XRD) (Rigaku MiniFlex 600) using the Cu Ka radiation (1
crease TiO2 photocatalyst efficiency has been carried out. The use of a nm ¼ 0.15418). The magnetic properties of nanocomposites were
magnetic core is due to its biocompatibility, low cost, and abundance for measured using a Vibrating Sample Magnetometer (VSM) (Lake Shore
easy catalyst recovery has been carried out that simply exits using a 7403) at an ambient temperature. The morphology and structure of
magnetic field (Singh et al., 2019). To reduce TiO2 bandgap and its use resultant composite nanoparticles was obtained from Field Emission
in the visible light, noble metals such as Ag, Pt, Pd and Au doped onto Scanning Electron Microscopy (FESEM) (TE-SCAN, MIRA3 FESEM
TiO2 can be used that simply is via coating of these metal ions on the Model). The elemental analysis was determined using an alternating
surface of a semiconductor (Bokare et al., 2013). This operation causes Electron Dispersion X-Ray Spectroscopy (EDS along with FESEM). Zeta
TiO2 photoresponse improvement and photodegradation under solar potential of TiO2@ZnFe2O4/Pd composite combination was measured
light irradiation occurs with an ease. In addition, it prevents the via Zetasizer 3000HS. The palladium distinctive features of Pd sup­
recombination of electrons and holes that are caused as the result of ported TiO2@ZnFe2O4 sample was determined via Inductively Coupled
photodegradation (Huang et al., 2015). Also, Due to its high reactivity, Plasma Optical Emission Spectrometry (ICP-OES, Leeman Direct
this nanoparticles have good adsorption ability due to their specific Reading Echelle). The Fourier-Transform Infrared Spectroscopy (Shi­
surface area. The nanoparticles are also used to correct and absorb madzo, FT_IR1650 Spectrophotometer, Japan) with KBr pellets was used
environmental pollution (Arsiya et al., 2017). Thus, the aim of this study for chemical nature display of the prepared composites at the range of
was magnetic nanocomposite synthesis of palladium supported 400–4000 cm 1. The reflexive spectrum of UV radiation i.e. the UV–Vis
TiO2@ZnFe2O4 with photodegradation ability of DCF under direct solar (DRS) spectra of the samples were recorded on the UV–Vis Spectro­
irradiation in the batch and continuous systems. photometer (Shimadzu, UV-2550, Japan). The Thermographic analysis
(TGA) was determined using the PerkinElmer System (USA) at 50� to
2. Experimental 800 � C in N atom.

2.1. Material
2.4. Batch system

All the reagents were purchased from Aldrich and Merck and used
Optimum pH levels, concentration of DCF and magnetic catalyst was
without further purification. The chemicals ZnCl2 (<98%), FeCl3.6H2O
determined using the batch system. In this study, the photocatalytic
(<97%), CH3COONa.3H2O (NaOAc <99%), NaOH (96%), HCl (97%),
system was subjected to direct solar light irradiation with an intensity of
ethylene glycol (EG), tetra butyl orthotitanate (TBOT) besides ethanol
72–76 Klux. A magnetic stirrer with 70 rotations/min was used for
and methanol were procured from Merck, Germany and Pd(NO3)22H2O
mixture of the samples. To adjust pH of the samples 0.1 M HCl and NaOH
from Aldrich Company. Moreover, DCF was purchased from Sigma-
were used. Before turning on the lamps and or exposure of the sample to
Aldrich Co. In all the stages, deionized water was used to prepare the
solar light irradiation, initially the solution was stirred for 30 min in
solutions.
dark, to assure the sorption – desorption equilibrium between the pho­
tocatalyst and Diclofenac. In a specified timescale, an amount of re­
2.2. Preparation of ZnFe2O4, TiO2@ZnFe2O4 and Pd supported action’s solution was withdrawn and after centrifugation and for
TiO2@ZnFe2O4 magnetic nanoparticles removal of residual particles a filter paper was used and finally
measured via spectrophotometer at the range of 276 nm. DCF photo­
ZnFe2O4 magnetic nanoparticles were synthesized via Solvothermal degradation efficiency was calculated as under (Ahmadpour et al.,
method based on the previous reports (Guo et al., 2016;Gu et al., 2016) 2019):
and the accurate details is mentioned as follows: 1.08 g FeCl3.6H2O (2 � �
mmol) and 0.82 g NaOAc (4 mmol) was dissolved in the ethylene glycol C t C0
Removal DCFð%Þ ¼ �100 (1)
(EG) mixture and stirred at RT for 30 min. The resultant brownish dark Ct
yellow color mixture was autoclaved for 4 h at 180 � C. After autoclaving
the solution at cold room temperature, the resultant black color sedi­ where C0 and Ct were the initial and final concentrations respectively
ment was separated from the mixture with a strong magnet and after after different time intervals of solar light irradiation.
washing with deionized water (DW) and ethanol, was dried in an oven at
70 � C for 24 h. 2.5. Continuous system
For preparation of TiO2@ZnFe2O4, 60 g of ZnFe2O4 nanoparticles
were mixed with (0.24 ml) DW and (60 ml) ethanol and was sonicated To assess the effect of column length and flow rate in breakthrough
for 20 min. In the next step, 200 ml of Tetrabutyl Orthotitanate (TBOT) curves, the system length (5, 10, 15 and 25 cm) and three flow rates (5,
and 10 ml ethanol under vigorous stirring (3 h at 40 � C) was added 10 and 15 mm/min) was used in the continuous system. In this study, 10
dropwise to the ferrite solution. At the reaction end, the resultant mg/l of DCF and 0.03 g/l of Pd supported TiO2@aZnFe2O4 composite
nanoparticle was separated with a magnet and washed with DW and catalyst was added to the system and subjected to direct solar light
ethanol and dried at 70 � C (Chi et al., 2013). irradiation. It is imperative to mention that in this system the entire
In the next stage, TiO2@ZnFe2O4//Pd nanoparticles were optimized parameters in the batch system were used. To adjust the flow

2
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

rate a peristaltic pump was used. Furthermore, a magnetic stirrer (with TiO2@ZnFe2O4 core/shell, the peaks at 2θ showed 25.75⁰, 27.2⁰, 35.8⁰,
70 rotations/min) to mix the sample and an aeration pump (1.5 l/min) 47.4⁰, 57.35⁰ and 62.95⁰ that pertain to surfaces (101), (110), (311),
to provide the required solution oxygen was used. In these experiments, (102), (422) and (511) of TiO2 respectively (JCPDS No. 1272-21)
the temperature was kept constant at 25 � C. The output liquid volume (Hasanpour et al., 2012). TiO2 peaks with tetrahedron shell anatas
(Veff) was achieved from the equation (Sayadi et al., 2019) structure in some of the peaks had ZnFe2O4 nuclei. The dispersion ra­
diation intensity reduction of ZnFe2O4 was resultant of TiO2 coating due
Veff ¼ Qt (2)
to absorption of X-ray via TiO2 shell. The peaks at 2θ, 46.5⁰ and 68.1⁰
pertained to the cubic surfaces of palladium (200) and (220) respec­
where Q was flow (ml/min) and t was total flow time (min).
tively (JCPDS No. 89-4897) and showed that metallic palladium exists in
the composition of composite material.
2.6. Kinetic degradation of DCF
3.3. Fourier transform infrared (FTIR) analysis
In general, the first order kinetic is used for the micropollutants
(Bahnemann et al., 2007). The k constant rate was calculated from the
FTIR spectra of ZnFe2O4, TiO2@ZnFe2O4 and TiO2@ZnFe2O4/Pd
first order equation:
nanoparticles are indicated at the range of 400–4000 cm 1 (see sup­
dC plementary material SM 1). From FTIR spectra, it was clear that ZnFe2O4
¼ kC (3)
dt was strong absorption group at 540 cm 1 level that is in accordance with
Fe–O vibrations (Rahimi et al., 2015). Moreover, Zn–O stretching vi­
where C is DCF concentration and k is reaction constant rate. The pho­ brations was depicted at 552 cm 1 and the band with 1384 cm 1 range
tocatalytic kinetic reactions can be described using the first order re­ could be observed with Ti–O–Ti vibrations in TiO2 (Dhiman et al.,
action for 10 mg/l concentration of DCF solutions. The constant rate (k) 2012). The other two peaks at 3250 cm 1 and 1644 cm 1 respectively
and correlation coefficient (R2) was assessed via the linear regression were created at the surface that pertained to O–H vibrations in the hy­
curve Ln (C0/C) and UV light irradiation time. The first order equations droxyl group absorbed from the water (Ullattil and Periyat, 2016). After
are as follows (Sharma et al., 2012): modification of TiO2@ZnFe2O4 with Pd the peaks appeared at 1627
C0 cm 1 and 2929 cm 1 respectively that show the presence of palladium
Ln ¼ Kt (4) in the composition.
C

here “k” is equation constant rate (min 1); C and C0 are the final and 3.4. X-ray photoelectron spectroscopy analysis
initial concentrations of DCF (mg/l) respectively.
X-ray photoelectron spectroscopy (XPS) is a powerful instrument to
3. Results and discussion study the elemental composition of the surface, the oxidation state, and
the binding energy of the core electron of the metal. The peaks related to
3.1. Characterization of TiO2@ZnFe2O4 nanoparticles O 1s, Fe 2p, Zn 2p, Ti 2p and palladium are clearly found in this spec­
trum. It contains two dominant band peaks corresponding to Fe2p3/2
3.1.1. Morphology and Fe2p1/2 at 711.6 and 724.2 eV, respectively. The Zn 2p XPS spectra,
The FESEM images showed that ZnFe2O4 and TiO2@ZnFe2O4 nano­ confirmed by the binding energies of Zn 2p3/2 and Zn 2p1/2 at 1020.3
particles were almost uniform in size and shape (see supplementary and 1043.2 eV, respectively. The Ti 2p, the two peaks 2p1/2 and 2p3/2 of
material SM 1). When TiO2@ZnFe2O4 surface is coated with Pd, the the Ti4þ were found in 464 eV and 458.3 eV, respectively. The two peaks
coarseness and size of particles increases. As observed, TiO2@ZnFe2O4/ at 530.7 and 529 eV are attributed to oxygen in ZnFe2O4 and Ti–O
Pd particles were 33.44–37.16 nm in size that are in concurrence with bonding to O, respectively (Wang et al., 2013). There is an XPS spectrum
XRD results. One of the reasons for compactness of agglomeration in of Pd binding energy with two peaks of 337.8 and 343.1 eV, which is
TiO2@ZnFe2O4/Pd nanoparticles is related to the magnetic properties of attributed to Pd 3d3/2 and Pd d35/2 (see supplementary material SM 2
ferrite nanoparticles (Nguyen et al., 2019). The absence of sonication for further details).
prior to imaging could also be the cause of nanoparticles compactness
(Moreno-Valencia et al., 2017; Wang et al., 2017). The elemental anal­ 3.5. UV-VIS spectrum (DRS) analysis
ysis for chemical purity of TiO2@ZnFe2O4/Pd sample was carried out
with X-ray Spectroscopy analysis (EDS). EDS spectrum showed that the For photoproperties perception of ZnFe2O4, TiO2@ZnFe2O4 and
elements Zn, Fe, O, Ti and Pd are present in the sample. Therefore, EDS TiO2@ZnFe2O4/Pd composite nanoparticles, UV–Vis spectrum (DRS) of
spectrum validates that Pd particles are present on the titanium mag­ the samples was carried out (Fig. 1a). As indicated, ZnFe2O4,
netic nanoparticles (see supplementary material SM 1). The reason for TiO2@ZnFe2O4 and TiO2@ZnFe2O4/Pd composite nanoparticles were
high TiO2 rate is that TiO2 nanoparticles lack magnetic interaction thus able to absorb the visible light. ZnFe2O4 nanoparticles can totally absorb
in nanoparticles washing phase, it is unable to separate them via the UV and visible light due to narrow bandgap. Therefore, it enhances the
magnet. electron-hole recombination (Yu, 2004). After TiO2 deposition, the ab­
sorbency of TiO2@ZnFe2O4 increases in visible light which is due to
3.2. Crystalline structure, X-ray diffraction (XRD) analysis synergy effects of TiO2 transfer and ZnFe2O4 absorption (Yu et al.,
2011). Later with palladium coating the absorbency in visible light
The structure of magnetic nanoparticles was determined using the X- significantly increases that is due to the surface plasmon resonance ef­
ray diffraction (XRD). The XRD pattern of ZnFe2O4 is completely allo­ fects of Pd arising from the visible light (Guo et al., 2013; Hasanpour
cated to the cubic phase of ZnFe2O4 (see supplementary material SM 1). et al., 2012; Rahimi et al., 2015; Dhiman et al., 2012; Ullattil and Per­
The diffraction peaks at ¼ 2θ showed 30.45� ,32.15� , 35.8� , 43.45� , iyat, 2016; Wang et al., 2013; Yu, 2004; Yu et al., 2011; Lin et al., 2012).
53.8� , 57.35� ,62.95� , 73.5� pertain to ZnFe2O4 and pertain to the crys­
tals surface (220), (222), (311), (400), (331), (422), (511) and (533) 3.5.1. Measurement of magnetic properties (VSM)
respectively (JCPDS 22-1012) (Chen et al., 2015; Yang et al., 2015; Guo The magnetic properties of nanoparticles were determined via a
et al., 2013). In case of TiO2@ZnFe2O4 composite, the cubic structure magnetometer (VSM). As shown in Fig. 1b all of them indicated high
peaks in the axis showed the composites; therefore it was clear that saturation magnet. Magnetic saturation level (Ms) of ZnFe2O4 nano­
ZnFe2O4 nanocrystals do not alter their phases. Besides XRD pattern of particles was 42.57 emu/g. After modification with TiO2 and doped Pd,

3
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

Fig. 1. UV–Vis spectrum (DRS) (a), and (b) VSM of ZnFe2O4, TiO2@ZnFe2O4 and TiO2@ZnFe2O4/Pd magnetic nanocomposite.

the magnetism saturation reached 37.32 and 27.28 emu/g respectively. 3.6. Thermodynamic stability
In fact, it can be stated that saturation magnetism of magnetic nano­
particles gradually reduced due to reduction of ZnFe2O4 content in the The thermal stability of TiO2@ZnFe2O4/Pd nanocatalysts and as well
composite nanoparticles. Although, TiO2@ZnFe2O4/Pd in relation to the percent performance of the functional groups on the magnetic sur­
ZnFe2O4 encounters a lower magnetism, but still is located in the face of nanoparticles was assessed using TGA analysis. The weight of all
paramagnetic domain wherein the resultant nanocomposite nano­ the samples at the temperature over 500 � C was stable. 8.93% weight
particles can be immediately separated from the solution under an loss in the samples occurred at 50⁰ to 500 � C wherein it guarantees the
external magnetic field. thermal stability of ZnFe2O4 nanoparticles at high temperature (see
supplementary material SM 3 for further details). After the ferrite
3.5.2. N2 adsorption-desorption isotherms coating with titanium dioxide in the initial stages weight loss was
To examine the specific surface area and the pore size distribution, observed. This was due to physical evaporation of alcohol and absorbed
N2 adsorption-desorption isotherms are performed. As Figure (2 a,b) water and loss of some of the residual organic solvents. After nano­
shows, the N2 adsorption-adsorption isotherms correspond to the IUPAC particles coating with palladium the thermal stability showed 31.7%
type III, which indicated little pore on the ZnFe2O4 surface morphology. reduction in the sample weight in TiO2@ZnFe2O4/Pd which is created in
The total pore volume and the surface area were 0.457 cm3 g 1 and the nanoparticles heating at 500 � C and it was due to degradation of
46.03 m2 g 1 respectively. The isotherm of the composite nanoparticles some of the organic molecules of solvent groups or hydroxyl present in
TiO2 @ ZnFe2O4 and TiO2 @ ZnFe2O4/Pd corresponds to the IUPAC IV the sample surface (Ba-Abbad et al., 2012) and as well dehydration of
and with H2 hysteresis loops, which indicates the mesoporous charac­ the water and phase structure of the conversion (Tekmen et al., 2008).
teristic (Chen et al., 2015). The average pore diameter of these nano­ At 800 � C and higher the sample became stable.
particles were 3.7 and 5/7 nm. The total pore volume and surface area of
these nanoparticles were 0.442 cm3g 1 and 349.25 m3/g 1, respec­
3.7. Zeta potential
tively. After coating with Pd metal nanoparticles, the surface area
increased to 131.18 m2 g 1 and the average pore diameter decreased to
Zeta potential has a role in the behavior of nanoparticles and its ef­
6.821 nm and the total pore volume 0.399cm3g-1. It should be noted
fect on the stability and durability rate of very tiny particles dispersed in
that, the mesoporous structure is suitable for enhancing specific surface
the aqueous solutions. Zeta potential determines the repulsive rate of
area and improving catalytic performance (Zhao et al., 2019).
similar and adjacent loaded particles dispersed in a solution. The par­
ticles which are adequately small, for higher stability need a higher Zeta
potential. The particles with low Zeta potential show an inclination to

Fig. 2. (a) N2 adsorption-desorption isotherms; (b) pore size distribution curve.

4
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

turn into mass and accumulate (Doostmohammadia et al., 2001). The material SM 4 for further details). As shown in Fig. 3a, the removal ef­
particles with zeta potential higher than �30 mV are stable due to ficiency after 120 min irradiation was 81.11, 73.48, 65.98 and 46.21%
electrostatic divergence between the particles that prevent accumula­ at pH’s 4, 5, 7 and 9 respectively. As observed, pH variations were
tion and finally are deposited (Chalasani and Vasudevan, 2013; Pierce effective in DCF removal efficiency wherein with pH reduction the
and Zhao, 2011). pHpzc is an important and effective factor on the ab­ removal efficiency increases and eventually at pH 4 the highest removal
sorbent’s load, and for better understanding, pH on the absorbent was efficiency was observed. In other words, pH is one of the effective factors
assessed, pH wherein the positive and negative loads of absorbent are in degradation rate of organic materials in a photocatalytic process and
equivalent and or in other words the absorbent has a neutral load is pH variations can affect DCF molecules adsorption on TiO2@Zn­
called an isoelectric point. In a manner that in pH < pHpzc, the absor­ Fe2O4/Pd catalyst surface. pH value affects the absorption capacity and
bent load due to presence of hydrogen (Hþ) ions gets a positive load and compounds dissociation, charge distribution on the catalyst’s surface
in pH > pHpzc, the absorbent load due to presence of hydroxide ions and oxidation potential of catalyst’s valence band. Therefore, pH vari­
(OH ) becomes negative (Kakavandi et al., 2016). The isoelectric point ations can affect DCF molecules absorption on TiO2@ZnFe2O4/Pd
of synthesized ZnFe2O4 nanoparticles was approximately equivalent to catalyst surfaces. pH ZPC is pH value wherein the load flow on absorption
6.3 which showed a positive load surface in ferrite. As observed, at surface is zero. The zero potential point for TiO2@ZnFe2O4/Pd catalyst
acidic pH, most of the DCF species have positive and negative load, is equivalent to 8.2 indicating that catalyst surface is activated in acidic
wherein the catalyst surface has a positive load; considering the low conditions and pH below 8.2 (Zarei et al., 2010). On the other hand, OH
surfaces in zinc ferrite it encounters a lower stability during the time radicals can be formed in the reaction between hydroxyl ions and pos­
period, whereas with TiO2 coating considering the repulsive energy itive holes wherein these positive holes are formed at low pHs. DCF
among the particles that signifies a negative energy between the parti­ degradation rate due to pH reduction and absorption load reduction
cles there is a higher stability and homogeneity. Therefore, with stability reduces at pH < PH ZPC. Therefore, the maximum DCF reduction
increase in the suspension, the absorption and removal of different occurred at pH ¼ 4. The probable description of this finding is ampho­
pollutant molecules on TiO2 surfaces increases (Wang and Ku, 2007; Xu teric behavior of semiconductor materials and TiO2@ZnFe2O4/Pd
et al., 2007). Besides, with increase of electrostatic gravity between the nanocomposite’s variation in surface charge properties that is propor­
DCF molecules and positive surface of catalytic nanoparticles, the ab­ tional with pH variations. This theory could be related to the presence of
sorption efficiency of drug increases (see supplementary material SM 3 sole electron pair on DCF molecules that causes electrostatic absorption
for further details). between negative DCF molecule and positive TiO2@ZnFe2O4/Pd com­
posite in the acidic conditions and vice-versa. The other reason for DCF
3.8. UV-light induced photodecomposition of DCF degradation efficiency increase at acidic pHs is that in acidic solutions
the high concentration of Hþ ion causes (OH ) hydroxyl radical for­
3.8.1. pH effect mation whereby via oxygen’s present in the solution, form superoxide
Initial pH is one of the most effective parameters on the photo­ radical and eventually converts to hydroxyl radical. The high efficiency
catalytic process affecting sorption/desorption capacity of target of this process at acidic pHs can be related to the following reactions in
organic compositions that in turn affects the peculiarities of photo­ the acidic conditions.
catalyst’s superficial charge and determines ionization state of catalyst’s 2HO2 ​ → O2∘ ​ þ ​ H2O2 (5)
surface (Safari et al., 2014). In fact, solution pH is in correlation with the
photocatalytic performance and shows that organic wastewater com­ H2O2 ​ þ ​ O2∘→OH∘ þ OH ​ þ ​ O2 (6)
positions differ in hydrophobity, dissolubility and varied behaviors. The
magnetic TiO2@ZnFe2O4/Pd nanocomposite was assessed for DCF e ​ þ ​ O2→O2∘ (7)
degradation at different pH’s (4, 5, 7 and 9). In this experiment, 10 mg/l
of initial drug concentration was mixed in 100 ml of aqueous solutions O2 ∘ ​ þ ​ H þ →2HO∘2 (8)
wherein for pH adjustment 0.1 M NaOH and HCl were used. Later,
Furthermore, based on the previous reports DCF exists more in ionic
0.005 g/l TiO2@ZnFe2O4/Pd composite catalyst was added to the
rather than molecular form in an alkaline condition. In a photocatalytic
considered solution. It mandates to mention that prior to sample’s
process, Diclofenac’s molecular form is more considered since molecular
exposure to solar light irradiation, initially the solution was stirred for
absorption of DCF on the catalyst surface is higher than its ionic ab­
30 min in dark, to assure the sorption/desorption equilibrium between
sorption (Arsiya et al., 2017). Therefore, formation of hydroxyl radicals
the photocatalyst and DCF. The photocatalytic degradation experiment
in acidic conditions is more which in turn causes degradation increase
after DCF absorption equilibrium was carried out on TiO2@ZnFe2O4/Pd
(Bagal and Gogate, 2014). Mugunthan et al., 2018, in their study on
composite surface with solar light irradiation. Fig. 3a depicts DCF
photocatalytic degradation of DCF under visible light using WO3/TiO2
removal at different time periods with varied pH’s (see supplementary

Fig. 3. The C/Co curve for the photodegradation of DCF on (a) pH; (b)catalyst dose; (c) DCF concentration.

5
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

nanoparticles, reported that the highest degradation occurred at acidic surface in the degradation process (Kermani et al., 2013). In addition,
condition due to Diclofenac’s molecular form (Meng et al., 2018). the removal efficiency can increase with concentration increase, but
On the other hand, it was determined that pH directly affects radical after a time period it reaches an equilibrium state because initially the
generation (Norzaee et al., 2017). One of the influential parameters in active sites on catalyst surface are easily accessible and after the passage
pH included the greater presence of Hþ ions in the acidic media, fol­ of time and or with concentration increase, these active sites are occu­
lowed by the formation of OH� radicals, as well as electrostatic reactions pied and efficiency reduces (Chen et al., 2015). Therefore, the reduction
between pollutants and catalyst surface (Busca et al., 2008). The reason of hydroxyl radicals’ formation can occur to an extent that affects the
for the increase in removal efficiency in acidic pH is that the catalyst photocatalytic activity of the catalyst (Muruganandham et al., 2006; Das
surface has a positive charge and causes more DCF degradation, as well et al., 2015).
as more radical hydroxyl production and decomposition in the acidic
media, thus increasing the removal efficiency. While in an alkaline 3.8.4. Disentangle of significant parameters using analysis of variance
condition, the surface of nanoparticles has a negative surface charge and The analysis of variance for removal efficiency of DCF using Pd-
reduced the production of radical hydroxyl and reduced the efficiency in coated TiO2@ZnFe2O4 magnetic photocatalyst was performed and the
the alkaline media (Kansal et al., 2010; Hosseinzadeh et al., 2012). Also, significant factors can extricate according to Pvalue. Hence, the pa­
reducing the efficacy of removal in alkaline conditions can be attributed rameters like pH, catalyst dose, DCF concentration, and aeration and
to the formation of insoluble compounds that reduced light intensity and magnetic stirrer have an effect on the photodegradiation of DCF,
reduced radical hydroxyl production (Dehghani et al., 2014). In addi­ whereas low Pvalue (Pvalue ¼ 0.00) obtained for all of the parameters
tion, it has been reported that changing the pH around the pHZPC causes and assigned as significant factors (Ekrami-Kakhki et al., 2018).
catalyst particles to adhere and thus settle the particles. This settling will Therefore, F value can demonstrate the importance of the parameters
reduce the efficiency of the degradation (Malato et al., 2009). (Abbasi et al., 2019). Consequently, the sequence importance of the
parameters is equal to catalyst dose (30.717) > DCF concentration
3.8.2. Catalyst dose effect (21.506)> pH (19.770) > aeration and magnetic stirrer (7.905).
Different catalyst concentrations are one of the important parame­
ters than can affect the degradation rate of DCF. As shown in Fig. 3b,
3.9. Kinetic photocatalytic degradation of DCF
DCF degradation percent and as well (C/C0) degradation ratio and time
intervals, the effect of photocatalyst concentration on DCF removal ef­
To assess kinetic photocatalytic degradation of DCF in the photo­
ficiency using TiO2@ZnFe2O4/Pd composite nanoparticles was evalu­
catalytic process at varied concentrations of the pollutant, the experi­
ated at the range of 0.005–0.1 g/l at the initial concentration of 10 mg/l
ments were conducted at optimal conditions (nanocatalyst dose: 0.03 g/
DCF and optimal pH ¼ 4 (see supplementary material SM 4 for further
l, contact time: 120 min and pH 4 at different DCF concentrations) to
details). It mandates to mention that prior to solar light irradiation
evaluate kinetic degradation of DCF. For conduction of this experiment,
exposure, initially the solution was stirred in dark for 30 min, to assure
initially the solution was kept in dark for 30 min and its photo­
sorption/desorption equilibrium between the photocatalyst and Diclo­
degradation and kinetic mechanism was determined (see supplementary
fenac. Initially the photolysis experiments showed that in case of lack of
material SM 4 for further details). As results reveal DCF photocatalytic
photocatalyst use the degradation rate reduces to an extent. This sig­
degradation follows the kinetic pseudo order one and can be well
nifies that light has an insignificant effect on DCF degradation. As
described via Langmuir-Hinshelwood model (Daghrir and Drogui, 2013;
observed, with addition of a photocatalyst to the initial degradation
Subramonian and Wu, 2014). As observed with pollutant concentration
process, considering the catalyst concentration, it increases up to a
increase the constant rate (k) reduces in a manner that it changes from
specific level and after that, removal occurred at a lower rate which
0.0172 min 1 at 10 mg/l DCF concentration to 0.0128 min 1 at 30 mg/l
signified the saturated photon absorption (Mugunthan et al., 2018). At
DCF concentration. The constant speed for different reaction concen­
low catalyst concentration, light transfers at a higher rate to the solu­
trations is tabulated in Table 1.
tion. The absorption dose increase from 0.005 to 0.1 g/l causes DCF
degradation increase since with absorbent’s dose increase the access and
active absorption sites between the absorbent and absorptive in a re­ 3.10. Comparison of PCP degradation rates by the synthesized catalysts
action increases. Even the other studies proved that catalyst dose in­
crease at specified levels; causes increase in the photocatalytic function The photocatalytic activities of ZnFe2O4, TiO2@ZnFe2O4 and
and surface access increase (Parra et al., 2002; Elmolla and Chaudhuri, TiO2@ZnFe2O4/Pd was evaluated with optimal DCF concentration (10
2010). In addition, more active radicals are formed. In some researches, mg/l) and pH ¼ 4 for DCF photodegradation in the aqueous solution
the accumulative effect of photocatalysts and light penetration lack in exposed to solar light irradiation for 120 min (see supplementary ma­
the solution has been stressed. In this study, catalyst rate increase caused terial SM 4 for further details). Prior to solar light irradiation exposure,
solution turbidity and lack of light penetration to the solution and initially the solution was kept in dark for 30 min. DCF absorption on
degradation efficiency reduction at the higher doses. catalyst’s surface in dark was very insignificant. When it was exposed to
light considering the catalyst type, DCF degradation increased. As
3.8.3. The initial DCF concentration effect observed, TiO2@ZnFe2O4/Pd magnetic composite encounters a higher
Photocatalytic degradation rate is intensely affected by the initial removal in relation to TiO2@ZnFe2O4 and ZnFe2O4 in the similar
concentration of target pollutant. To assess the initial pollutant con­ experimental conditions. This is due to bandgap energy related to ab­
centration effect on the photocatalytic efficiency, a wide range of DCF sorption in the visible region which can lead to efficiency improvement
concentration from 10 to 30 mg/l was prepared and the other experi­
mental conditions at initial pH 4 and catalyst concentration 0.03 g/l was Table 1
constant and results are depicted in Fig. 3c, and see supplementary Reaction rate constant of diclofenac photocatalytic degradation with different
material SM 4 for further details. As observed, with DCF concentration concentrations.
increase the removal efficiency reduces in a manner that the maximum Diclofenac concentration (mg/L) Equation K0(min 1
) R2
degradation rate occurred at 10 mg/l concentration. This finding can be 10 Y ¼ 0.0172x þ 1.5067 0.0172 0.9952
due to reason that at low concentration of pollutants, the probability of a 15 Y ¼ 0.0137x þ 1.2228 0.0137 0.9916
reaction between reactive species and pollutant increases and at higher 20 Y ¼ 0.0131x þ 1.0573 0.0131 0.9808
DCF concentration, the excessive amount of target molecules are accu­ 25 Y ¼ 0.013x þ 0.7941 0.013 0.9812
30 Y ¼ 0.0128x þ 0.5457 0.0128 0.9958
mulated on the catalyst surface that causes reduction of an effective

6
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

in the catalytic degradation. Therefore, after subjection to solar light


e ​ þ ​ O2 →O2 (9)
irradiation for 120 min, DCF degradation efficiency via TiO2@ZnFe2O4/
Pd nanocomposite was 84.87% which was much higher than
H þ ​ þ ​ O2� →HO2� (10)
TiO2@ZnFe2O4 (73.33%) and ZnFe2O4 (49.9%). Therefore, TiO2@Zn­
Fe2O4 composite catalyst had the highest photocatalytic activity among 2HO2� → H2O2 ​ þ ​ O2 (11)
the different catalysts in this study exposed to solar light irradiation that
was due to synergy effect between Pd, TiO2 and ZnFe2O4 nanoparticles. H2 O2 →2OH� (12)
Likewise, palladium causes load and Plasmon Resonance Energy transfer
between TiO2. On this basis, considering the interactions, it eases the H2 O2 ​ þ ​ O2� →� OH ​ þ ​ OH ​ þ ​ O2 (13)
transfer and separation of photogenerated electrons and holes that leads
to higher improvement in the photocatalytic function. H2 O2 ​ þ ​ e →� OH ​ þ ​ OH (14)
In fact, airflow causes production of higher oxygen besides reaction
3.11. Aeration effect in DCF photodegradation process with excited electrons via photon and forms O2 anions. Formation of O 2
causes degradation rate increase in the aeration flow increase (Lee et al.,
One of the persisting issues in photocatalytic process is solution 2004). However, the flow over 1.5 l/m causes light penetration due to
immobilization, since the catalyst is positioned on an immobilized solid formation of large amount of bubbles during photocatalysis radiation.
carrier (Minsker et al., 2001). This limitation is resolved by complete
mixing and turbulence formation in the reactor. Likewise, the presence 3.12. Mechanism of DCF photodegradation process
of solution oxygen has an important role in the photocatalytic degra­
dation of organic materials. The oxygen molecule can as an electron Fig. 4 shows the possible reaction mechanisms for DCF photo­
acceptor get trapped and separate the electron from positive holes to degradation by Pd supported TiO2@ZnFe2O4 nanocomposites in the
assist the reduction of probable electron-hole recombination. Never­ presence of solar light. When palladium nanoparticles come in contact
theless, in this study, the simultaneous effect of aeration and magnetic with TiO2 nanoparticles without solar light radiation, due to the energy
stirrer on DCF photocatalytic degradation via TiO2@ZnFe2O4/Pd level of palladium is higher than TiO2, The electrons are transferred
nanocomposite at 120 min in three aeration conditions; use of magnetic from palladium to TiO2 to balance the two systems. Whenever the
stirrer and simultaneous use of aeration/magnetic stirrer was assessed TiO2@ZnFe2O4/Pd TiO2 photocatalysts are irradiated under solar light
and the results are depicted in supplementary material SM 5. As with photon energy that is higher than the TiO2 band gap, optical
observed, the highest degradation rate (86.5%) was achieved in the excited electrons and holes are formed in the conduction band (CB) and
aeration/magnetic stirrer system, whereas in the sole system of aeration the valence band (VB) of TiO2 (Nguyen and Doong, 2017; Ahmadpour
and magnetic stirrer the degradation rates were 81.73% and 78.02% et al., 2020). Since palladium nanoparticles have a high ability to
respectively. The aeration caused oxygen production in the solution that accumulate large amounts of electrons, the transfer of photoexcited
acts as an electron acceptor (Chong et al., 2010). The presence of elec­ electrons from TiO2 to palladium nanoparticles causes the energy level
tron acceptor causes increased separation of photogenerated electrons of the composites to shift to negative potential. Therefore the electrons
and holes in the photocatalytic process and eventually its improvement. in the CB can be transferred from TiO2 to Pd. Therefore, Pd nanoparticles
In fact, this system provides the oxygen required for photoelectrons act as electron acceptors, and results in reduced electron-hole recom­
evacuation and forms superoxide anionic radicals. The role of oxygen as bination and prolongs the lifetime of the electron-hole pairs. Afterward,
an electron acceptor is summarized as the following equations (Wang the electrons can be trapped by O2 and the holes can be captured trapped
et al., 2002). The molecular oxygen as an electron acceptor for trap and by the surface hydroxyl group, eventually leading to the formation of
removal of electrons uses TiO2@ZnFe2O4/Pd catalyst particles surface to superoxide anion radical (O2) and hydroxyl radical species (OH) (Kumar
reduce the free electron, the reaction of absorbed oxygen with photo­ and Devi, 2011). This radical groups (�O2 and �OH) are highly reactive
generated electrons at TiO2@ZnFe2O4/Pd catalyst surface is relatively to DCF degradation.
slow and could control the speed in photocatalytic oxidation reaction
(Wang et al., 2002).

Fig. 4. Schematic illustration of proposed photocatalytic mechanisms of ZnFe2O4@TiO2/Pd nanocomposite.

7
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

3.13. Photocatalytic degradation of DCF in higher volume the maximal removal rate of the target pollutants (Zhang et al., 2018).
The column length breakthrough curves for DCF removal are demon­
In this study, DCF removal was also evaluated at higher volume. The strated at different length of 5–25 cm. In this study, the initial concen­
experimental conditions were as follows: in a 1000 ml vessel containing tration of DCF was 10 mg/l and flow rate was 5 ml/min. As observed
DCF pollutant solution along with other optimized parameters (pH ¼ 4, with column length increase from 5 to 25 cm, the breakthrough curve
catalyst concentration 0.03 gm/l, drug dose 10 mg/l) was used (see slope reduces. This could be due to an access to pollutant connector sites
supplementary material SM 4 for further details). The present results in the higher column lengths for absorption and resultantly the mass
showed that with the passage of time the degradation efficiency in­ transfer zone is higher (Jain et al., 2013; Mishra et al., 2013). The
creases wherein this degradation rate under solar light irradiation after reaching time to the column break point at the length of 5, 10, 15 and 25
120 min was 78.52%. cm was 80, 65, 50 and 45 min respectively, and DCF degradation rate
reduced from 83.49% to 90.48% from 5 to 25 cm length. Besides, the
higher column length provides a longer residential phase of pollutant in
3.14. DCF removal in the photocatalytic process in the continuous system the column. Therefore, a higher volume of DCF pharmaceutical
pollutant can be filtered (see supplementary material SM 5 for further
3.14.1. Flow rate effect details).
Flow rate is an important parameter in the photocatalytic processes,
since it determines the pollutant’s contact time with an absorbent in the
column length. The breakthrough curves at three different flows (5, 10, 3.15. Reusability of the photocatalytic nanoparticles
15 ml/min) are revealed in supplementary material SM 5. As observed,
the progressive as well as saturation period reduces with flow rate in­ An important point for the application of photocatalytic nano­
crease since with higher flow the retention time in the column is not particles in the water filtration process was separation and reuse of
sufficiently high to create absorption equilibrium in the column. nanoparticles. For this reason, reuse of synthetized nanocomposite was
Therefore, considering the mass transfer limitations and sufficient assessed as a photocatalyst for DCF degradation during 5 recycles at
retention time, the degradation performance of DCF reduces (Bulgariu optimum conditions in a batch system. After each cycle, the nanocatalyst
and Bulgariu, 2016). In real time where the flow increases from 5 to 15 was separated via an external magnet from the solution and washed
ml/min, the retention time in the column reduces. Whereby, with flow several times with deionized water and ethanol and later dried in an
rate reduction due to higher retention time in the column and contact oven and used in the next stage of DCF removal (Shekari et al., 2017). As
increase, DCF degradation efficiency increases. demonstrated in Fig. 5a, TiO2@ZnFe2O4/Pd catalysts have relatively
stable photocatalytic activity and are recyclable and can be reused in a
3.14.2. Column length effect manner that after 5 photocatalyst recycles, DCF degradation reduced
Column length is an essential factor that affects the removal capacity from 86.1% in the first stage to 71.38% in the fifth stage. As denoted the
and estimation time of absorption bed. Therefore, a suitable column photocatalytic activity loss was insignificant which signifies the photo­
length and or moisture volume increases adsorbent’s access and assists catalytic stability of these nanoparticles during the photocatalytic

Fig. 5. (a) Efficiency of reuse on activity of TiO2@ZnFe2O4/Pd nanoparticles; (b) the XRD spectra, and (c) FTIR spectra of TiO2@ZnFe2O4/Pd before and after
recycling test.

8
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

reactions. Therefore, its application from the commercial viewpoint due Software, Methodology.
to its reuse potentiality is very useful. The Inductively Coupled Plasma
Atomic Emission Spectrometry (ICP-AES) showed that only a very little Acknowledgements
amount (less than 1%) of Pd metal was eliminated after five recycles of
magnetic catalyst. This paper was adapted from a Ph.D. student thesis (ID Num­
However, XRD and FTIR results of TiO2@ZnFe2O4/Pd photocatalyst ber:1156/1397) in the Department of Environmental Engineering, the
were used after five recycles. XRD results before and after degradation Faculty of Natural Resources and Environment, the University of Bir­
showed nil apparent change in the crystalline structure of nanoparticles jand, which has been conducted with the financial support the Iran
after five recycles (Fig. 5b). Besides, FTIR analysis of TiO2@ZnFe2O4/Pd National Science Foundation (INSF) (Grant No. 97024168/1398). The
nanocomposite was carried out before and after DCF degradation authors of this research highly appreciate the support of Faculty of
(Fig. 5c). Based on Fig. 5c, much variation was not depicted between the Natural Resources and Environment, University of Birjand.
two FTIR patterns. Considering Fig. 5c, the additional peaks of peak
1707 cm 1 pertained to C– – O vibrations, whereas peaks 3347 cm 1 and Appendix A. Supplementary data
1
1178 cm pertained to O–H and C–O stretching vibrations respectively
in the fifth reuse. As observed much variation did not occur in the FTIR Supplementary data to this article can be found online at https://doi.
chart peaks before and after the degradation process. This result showed org/10.1016/j.jenvman.2020.110964.
that TiO2@ZnFe2O4/Pd nanocomposite had a very good chemical sta­
bility and can safely be used as a reusable photocatalyst in most of the References
applications.
Abbasi, S., Hasanpour, M., Ahmadpoor, F., Sillanp€ a€
a, M., Dastan, D., Achour, A., 2019.
Application of the statistical analysis methodology for photodegradation of methyl
3.16. TOC analysis orange using a new nanocomposite containing modified TiO2 semiconductor with
SnO2. Int. J. Environ. Anal. Chem. 1–17. https://doi.org/10.1080/
TOC analysis was used to accurately assess the degradation rate of 03067319.2019.1662414.
Abellan, M.N., Bayarri, B., Gimenez, J., Costa, J., 2007. Photocatalytic degradation of
DCF molecules during photodegradation. TOC measure exhibits the sulfamethoxazole in aqueous suspension of TiO2. J. Appl. Catal. B-Environ. 74 (3–4),
carbon rate present in the organic combinations wherein it more accu­ 233–241.
rately reflects the total amount of organic combinations available in the Acu~na, V., Ginebreda, A., Mor, J.R., Petrovic, M., Sabater, S., Sumpter, J., Barcel� o, D.,
2015. Balancing the health benefits and environmental risks of pharmaceuticals:
solution. For TOC removal rate at optimal conditions in the continuous diclofenac as an example. J. Environ. Int. 85, 327–333.
system; the photocatalytic reaction was evaluated after 30 min darkness. Ahmadpour, N., Sayadi, M.H., Sobhani, S., Hajiani, M., 2020. A potential natural solar
TOC analysis results demonstrated (see supplementary material SM 6 for light active photocatalyst using magnetic ZnFe2O4 @ TiO2/Cu nanocomposite as a
high performance and recyclable platform for degradation of naproxen from aqueous
further details) that TOC degradation rate at 120 min was 39.20%. TOC solution. J. Clean. Prod. 268 (20), 122023 https://doi.org/10.1016/j.
increase in the initial hours of light irradiation was due to light ab­ jclepro.2020.122023.
sorption from DCF or formation of intermediates. Therefore, these pro­ Ahmadpour, N., Sayadi, M.H., Verma, A., Mansouri, B., 2019. Ultrasonic degradation of
ibuprofen from the aqueous solution in the presence of titanium dioxide
duced intermediates are not so sensitive in relation to photolysis. This
nanoparticles/hydrogen peroxide. J. Des. Water Treat 145, 291–299.
means that due to intermediates sensitivity to photolysis the degrada­ Arsiya, F., Sayadi, M.H., Sobhani, S., 2017. Arsenic (III) adsorption using palladium
tion was lower wherein DCF was degraded via photolysis process nanoparticles from aqueous solution. J. Water Environ. Nanotechnol. 2 (3),
(Abellan et al., 2007). 166–173.
Ba-Abbad, M.M., Kadhum, A.A.H., Mohamad, A.B., Takriff, M.S., Sopia, K., 2012.
Synthesis and catalytic activity of TiO2 nanoparticles for photochemical oxidation of
4. Conclusion concentrated chlorophenols under direct solar radiation. Int. J. Electrochem Sci. 7
(6), 4871–4888.
Bagal, M.V., Gogate, P.R., 2014. Degradation of diclofenac sodium using combined
TiO2@ZnFe2O4/Pd nanocomposites were prepared using the pho­ processes based on hydrodynamic cavitation and heterogeneous photocatalysis.
todeposition method and identified using XRD, VSM, FESEM/EDX, FTIR, J. Ultrasonics Sonochem. 21 (3), 1035–1043. https://doi.org/10.1016/j.
GTA, UV–Vis, Zeta Potential and ICP-OES techniques. The resultant ultsonch.2013.10.020.
Bahnemann, W., Muneer, M., Haque, M.M., 2007. Titanium dioxide-mediated
composite material had catalytic function and an excellent degradation photocatalysed degradation of few selected organic pollutants in aqueous
efficiency using the optimal parameters in the reduction of DCF suspensions. Catal. Today 124, 133–148.
pollutant and showed a super magnetic behavior that caused magnetic Bokare, A., Pai, M., Athawale, A.A., 2013. Surface modified Nd doped TiO2 nanoparticles
as photocatalysts in UV and solar light irradiation. Sol. Energy 91, 111–119.
separation and easy recovery of nanocatalysts from the reaction Bulgariu, D., Bulgariu, L., 2016. Potential use of alkaline treated algae waste biomass as
mixture. In addition, this composite catalyst has a thermal stability, high sustainable biosorbent for clean recovery of cadmium (II) from aqueous media: batch
photocatalytic activity, competent magnetic separation, good chemical and column studies. J. Clean. Prod. 112, 4525–4533.
Busca, G., Berardinelli, S., Resini, C., Arrighi, L., 2008. Technologies for the removal of
stability and relatively low cost which causes it to act as a green reliable
phenol from fluid streams: a short review of recent developments. J. Hazard Mater.
photocatalyst with superior photocatalytic activity to remove the pol­ 160, 265–288.
lutants under the solar light irradiation. Therefore, it can as a promising Calza, P., Sakkas, V.A., Medana, C., Baiocchi, C., Dimou, A., Pelizzetti, E., Albanis, T.,
user have an application in the high scale photocatalytic reactions in the 2006. Photocatalytic degradation study of diclofenac over aqueous TiO2 suspensions.
J. Appl. Catal. B: Environ. Times 67, 197–205.
industries. Chalasani, R., Vasudevan, S., 2013. Cyclodextrin-functionalized Fe3O4@ TiO2: reusable,
magnetic nanoparticles for photocatalytic degradation of endocrine-disrupting
Declaration of competing interest chemicals in water supplies. ACS Nano 7 (5), 4093–4104.
Chen, X., Burda, C., 2008. The electronic origin of the visible-light absorption properties
of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 130 (15), 5018–5019.
The authors declare that they have no known competing financial Chen, X., Dai, Y., Liu, T., Guo, J., Wang, X., Li, F., 2015. Magnetic core–shell carbon
interests or personal relationships that could have appeared to influence microspheres (CMSs)@ZnFe2O4/Ag3PO4 composite with enhanced photocatalytic
activity and stability under visible light irradiation. J. Mol. Catal. Chem. 409,
the work reported in this paper. 198–206.
Chi, Y., Yuan, Q., Li, Y., Zhao, L., Li, N., Li, X., Yan, W., 2013. Magnetically separable
CRediT authorship contribution statement Fe3O4@SiO2@ TiO2-Ag microspheres with well-designed nanostructure and
enhanced photocatalytic activity. J. Hazard Mater. 262, 404–411.
Chong, M.N., Jin, B., Chow, C.W.K., Saint, C., 2010. Recent developments in
Najmeh Ahmadpour: Writing - original draft, Data curation, photocatalytic water treatment technology: a review. J. Water Resour. 44,
Methodology. Mohammad Hossein Sayadi: Supervision, Validation, 2997–3027.
Writing - review & editing. Sara Sobhani: Visualization, Conceptuali­
zation, Investigation, Writing - review & editing. Mahmood Hajiani:

9
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

Czili, H., Horv�ath, A., 2009. Photodegradation of chloroacetic acids over bare and silver- Lin, X., Rong, F., Fu, D., Yuan, C., 2012. Enhanced photocatalytic activity of fluorine
deposited TiO2: identification of species attacking model compounds. J. A doped TiO2 by loaded with Ag for degradation of organic pollutants. Powder
mechanistic approach. Appl. Catal. B: Environ. Times 89, 342–348. Technol. 219, 173–178.
Daghrir, R., Drogui, P., 2013. Tetracycline antibiotics in the environment: a review. Malato, S., Fern� andez-Ibanez, P., Maldonado, M.I., Blanco, J., Gernjak, W., 2009.
J. Environ. Chem. Lett. 11 (3), 209–227. Decontamination and disinfection of water by solar photocatalysis: recent overview
Das, L., Barodia, S.K., Sengupta, S., Basu, J.K., 2015. Aqueous degradation kinetics of and trends. Catal. Today 147, 1–59.
pharmaceutical drug diclofenac by photocatalysis using nanostructured Meng, X., Zhuang, Y., Tang, H., Lu, C., 2018. Hierarchical structured ZnFe2O4@SiO2@
titania–zirconia composite catalyst. Int. J. Environ. Sci. Technol. 12, 317–326. TiO2 composite for enhanced visible-light photocatalytic activity. J. Alloys Compd.
https://doi.org/10.1007/s13762-013-0466-y. 761, 15–23. https://doi.org/10.1016/j.jallcom.2018.05.150.
De Oliveira, S.R., Barioni, L.G., Pellegrino, G.Q., Moran, D., 2018. The role of agricultural Minsker, K.S., Zakharov, V.P., Berlin, A.A., 2001. Plug flow tubular reactors: a new types
intensification in Brazil’s Nationally Determined Contribution on emissions of industrial apparatus. J.Theor. Found. Chem. Eng. 35 (2), 162–167.
mitigation. J. Agric. Syst. 161, 102–112. Mishra, V., Balomajumder, C., Agarwal, V.K., 2013. Adsorption of Cu (II) on the surface
Dehghani, M., Nasseri, S., Ahmadi, M., Samaei, M.R., Anushiravani, A., 2014. Removal of of nonconventional biomass: a study on forced convective mass transfer in packed
penicillin G from aqueous phase by Feþ 3-TiO2/UV-A process. J. Environ. Health Sci. bed column. J. Waste Manag. 1–8.
Eng. 12, 56–63. Moreno-Valencia, E.I., Paredes-Carrera, S.P., S� anchez-Ochoa, J.C., Flores-Valle, S.O.,
Dhiman, N., Singh, B.P., Gathania, A.K., 2012. Synthesis and characterization of dye- Avendano-G� omez, J.R., 2017. Diclofenac degradation by heterogeneous
doped TiO2-SiO2 core-shell composite microspheres. J. Nanophotonics 6, 063511- photocatalysis with Fe3O4/TixOy/activated carbon fiber composite synthesized by
1–063511-10. ultrasound irradiation. J. Mater. Res. Express 4, 115026. /10.1088/2053-1591/
Doostmohammadia, A., Monshia, A., Salehic, R., Fathia, M.H., Seyedjafarid, E., aa981f.
Shafie, A., Soleimani, M., 2001. Cytotoxicity evaluation of 63s bioactive glass and Mugunthan, E., Saidutta, M.B., Jagadeeshbabu, P.E., 2018. Visible light assisted
bone-derived hydroxyapatite particles using human bone-marrow stem cells. photocatalytic degradation of diclofenac using TiO2-WO3 mixed oxide catalysts.
J. Biomed Pap Med Fac Univ Palacky Olomouc Czech Repub 155 (4), 323–326. J. Environ. Nanotechnol. Monit. Manag. 10, 322–330. https://doi.org/10.1016/j.
https://doi.org/10.5507/bp.2011.028 doi. enmm.2018.07.012.
Ekrami-Kakhki, M.S., Abbas, S., Farzaneh, N., 2018. Statistical analysis of the Muruganandham, M., Shobana, N., Swaminathan, M., 2006. Optimization of solar
electrocatalytic activity of Pt nanoparticles supported on novel functionalized photocatalytic degradation conditions of Reactive Yellow 14 azo dye in aqueous
reduced graphene oxide-chitosan for methanol electrooxidation. Electron. Mater. TiO2. J. Mol. Catal. Chem. 246 (1–2), 154–161.
Lett. 14 (1), 70–78. Naddeo, V., Belgiorno, V., Kassinos, D., Mantzavinos, D., Meric, S., 2010. Ultrasonic
Elmolla, E.S., Chaudhuri, M., 2010. Degradation of amoxicillin, ampicillin and cloxacillin degradation, mineralization and detoxification of diclofenac in water: optimization
antibiotics in aqueous solution by the UV/ZnO photocatalytic process. J. Hazard. of operating parameters. Ultrason. Sonochem. 17 (1), 179–185. . doi.org/10.1016/j.
Mater. 173 (1–3), 445–449. ultsonch.2009.04.003.
Gu, W., Xie, Q., Qi, C., Zhao, L., Wu, D., 2016. Phosphate removal using zinc ferrite Nguyen, T.B., Doong, R.A., 2017. Heterostructured ZnFe2O4/TiO2 nanocomposites with
synthesized through a facile solvothermal technique. J. Powder Technol. 301, a highly recyclable visible-light-response for bisphenol a degradation. J. RSC Adv. 7,
723–729. 50006–50016.
Guo, J.F., Ma, B., Yin, A., Fan, K., Dai, W.L., 2013. Highly stable and efficient Ag/AgCl@ Nguyen, T.B., Huang, C.P., Doong, R., 2019. Photocatalytic degradation of bisphenol A
TiO2 photocatalyst: preparation, characterization, and application in the treatment over a ZnFe2O4/TiO2 nanocomposite under visible light. J. Sci. Total Environ. 646,
of aqueous hazardous pollutants. J. Hazard Mater. 211–212, 77–82. 745–756.
Guo, X., Lin, Y., Wan, J., Yu, X., 2016. Study on preparation of SnO2-TiO2/Nano-graphite Norzaee, S., Bazrafshan, E., Djahed, B., Kord Mostafapour, F., Khaksefidi, R., 2017. UV
composite anode and electro-catalytic degradation of ceftriaxone sodium. activation of persulfate for removal of penicillin G antibiotics in aqueous solution.
J. Chemosphere 164, 421–429. Sci. World J. 2017, 1–6.
Guo, R., Xia, X., Zhang, X., Li, B., Zhang, H., Cheng, X., Xie, M., Cheng, Q., 2018. Parra, S., Sarria, V., Malato, S., P�eringer, P., Pulgarin, C., 2002. Photochemical versus
Construction of Ag3PO4/TiO2 nano-tube arrays photoelectrode and its enhanced coupled photochemical–biological flow system for the treatment of two
visible light driven photocatalytic decomposition of diclofenac. J. Separ. Purif. biorecalcitrant herbicides: metobromuron and isoproturon. Appl. Catal., B 27,
Technol. 200, 44–50. 153–168.
Hasanpour, A., Niyaifar, M., Amighian, M.H.J., 2012. A novel non-thermal process of Perez-Estrada, L.A., Malato, S., Gernjak, W., Aguera, A., Thurman, E.M., Ferrer, I.,
TiO2- shell coating on Fe3O4-core nanoparticles. J. Phys. Chem. Solid. 73, Fernandez-Alba, A.R., 2005. Photo-fenton degradation of diclofenac: identification
1066–1070. of main intermediates and degradation pathway. J. Environ. Sci. Technol. 39,
Hasija, V., Raizada, P., Sudhaik, A., Sharma, K., Kumar, A., Singh, P., Thakur, V.K., 2019. 8300–8306.
Recent advances in noble metal free doped graphitic carbon nitride based Pierce, D.T., Zhao, J.X., 2011. Trace Analysis with Nanomaterials. Wiley -VCH Verlag
nanohybrids for photocatalysis of organic contaminants in water: a review. J. Appl. GmbH & Co. KGaA.
Mater. Today. 15, 494–524. Rahimi, R., Heidari-Golafzani, M., Rabbani, M., 2015. Preparation and photocatalytic
Hasija, V., Raizada, P., Sudhaik, A., Singh, P., Thakur, V.K., Khan, A.A., 2020. application of ZnFe2O4@ZnO core–shell nanostructures. J. Superlattices and
Fabrication of Ag/AgI/WO3 heterojunction anchored P and S co-doped graphitic Microstructures 85, 497–503.
carbon nitride as a dual Z scheme photocatalyst for efficient dye degradation. Solid Ravina, M., Campanella, L., Kiwi, J., 2002. Accelerated mineralization of the drug
State Sci. 1, 100:106095. diclofenac via fenton reactions in a concentric photo-reactor. J. Water Resour. 36
Hosseinzadeh, E., Zare, M., Torabi, E., Rahimi, S., Shokouhi, R., 2012. Sodium alginate (14), 3553–3560.
magnetic beads for removal of acid cyanine 5R from aqueous solution. Bimonth. J. Safari, G.H., Hoseini, M., Seyedsalehi, M., Kamani, H., Jaafari, J., Mahvi, A.H., 2014.
Hormozgan Univ. Med. Sci. 16, 101–111. Photocatalytic degradation of tetracycline using nanosized titanium dioxide in
Huang, J., Guo, X., Wang, B., Al, E., 2015. Synthesis and photocatalytic activity of Mo- aqueous solution. Int. J. Environ. Sci. Technol. 12 (2), 603–616. https://doi.org/
doped TiO2 nanoparticles. J. Spectrosc. 8. https://doi.org/10.1155/2015/681850. 10.1007/s13762-014-0706-9.
Article ID 681850). Sayadi, M.H., Sobhani, S., Shekari, H., 2019. Photocatalytic degradation of azithromycin
Jain, M., Garg, V.K., Kadirvelu, K., 2013. Cadmium (II) sorption and desorption in a fixed using GO@Fe3O4/ZnO/SnO2 nanocomposites. J. Clean. Prod. 232, 127–136. https://
bed column using sunflower waste carbon calcium-alginate beads. J. Bioresour. doi.org/10.1016/j.jclepro.2019.05.338.
Technol. 129, 242–248. Sharma, S., Mukhopadhyay, M., Murthy, Z., 2012. Rate parameter estimation for 4-
Kakavandi, B., Jonidi Jafari, A., Rezaei Kalantary, R., Nasseri, S., Esrafili, A., chlorophenol degradation by UV and organic oxidants. J. Ind. Eng. Chem. 18,
Gholizadeh, A., 2016. Simultaneous adsorption of lead and aniline onto magnetically 249–254.
recoverable carbon: optimization, modeling and mechanism. J. Chem. Technol. Sharma, S., Dutta, V., Singh, P., Raizada, P., Rahmani-Sani, A., Hosseini-
Biotechnol. 91 (12), 3000–3010. Bandegharaei, A., Thakur, V.K., 2019. Carbon quantum dot supported
Kansal, S.K., Ali, A.H., Kapoor, S., 2010. Photocatalytic decolorization of biebrich scarlet semiconductor photocatalysts for efficient degradation of organic pollutants in
dye in aqueous phase using different nanophotocatalysts. Desalination 259, water: a review. J. Clean. Prod. 228, 755–769.
147–155. Shekari, H., Sayadi, M.H., Rezaei, M.R., Allahresani, A., 2017. Synthesis of nickel ferrite/
Kermani, M., Bahrami Asl, F., Farzadkia, M., Esrafili, A., Salahshur Arian, S., titanium oxide magnetic nanocomposite and its use to remove hexavalent chromium
Arfaeinia, H., 2013. Degradation efficiency and kinetic study of metronidazole by from aqueous solutions. J. Surf. Interfaces 8, 199–205.
catalytic ozonation process in presence of mgo nanoparticles. J. Urmia Univ. Med. Singh, P., Sharma, K., Hasija, V., Sharma, V., Sharma, S., Raizada, P., Thakur, V.K., 2019.
Sci. 24 (10), 839–850. Systematic review on applicability of magnetic iron oxides–integrated photocatalysts
Kim, I., Tanaka, H., 2009. Photodegradation characteristics of PPCPs in water with UV for degradation of organic pollutants in water. Mater. Today Chem. 14, 1–27.
treatment. J. Environ. Int. 5, 793–802. Subramonian, W., Wu, T.Y., 2014. Effect of enhancers and inhibitors on photocatalytic
Kumar, S.G., Devi, L.G., 2011. Review on modified TiO2 photocatalysis under UV/visible sunlight treatment of methylene blue. J. Water Air Soil Pollut. 225, 1922.
light: selected results and related mechanisms on interfacial charge carrier transfer Tekmen, C., Suslu, A., Cocen, U., 2008. Titania nanofibers prepared by electrospinning.
dynamics. J. Phys. Chem. 115, 13211–13241. J. Mater. Lett. 62, 4470–4472.
Lee, J.M., Kim, M.S., Kim, B., 2004. Photodegradation of bisphenol-A with TiO2 Ullattil, S.G., Periyat, P., 2016. A {’} one pot {’} gel combustion strategy towardsTiþ3 self-
immobilized on the glass tubes including the UV light lamps. J. Water Resour. 38, doped {’}black{’} anatase TiO2-x solar photocatalyst. J. Mater. Chem. 4 (16),
3605–3613 [CrossRef] [PubMed]. 5854–5858. https://doi.org/10.1039/C6TA01993.
Lim, N.R., Shohayeb, B., Zaytseva, O., Mitchell, N., Millard, S.S., Ng, D.C.H., Quinn, L.M., Vieno, N.M., Harkki, H., Tuhkanen, T., Kronberg, L., 2007. Occurrence of
2017. Glial-specific functions of microcephaly protein WDR62 and interaction with pharmaceuticals in river water and their elimination in a pilot-scale drinking water
the mitotic kinase AURKA are essential for Drosophila brain growth. J. Stem Cell treatment plant. J. Environ. Sci. Technol. 41, 5077–5084.
Rep. 9 (1), 32–41.

10
N. Ahmadpour et al. Journal of Environmental Management 271 (2020) 110964

Vogna, D., Marotta, R., Napolitana, A., Andreozzi, R., d’Ischia, M., 2006. Advanced Yu, Y.T., 2004. Preparation of nanocrystalline TiO2-coated coal fly ash and effect of iron
oxidation of the pharmaceutical drug of diclofenac with UV/H2O2 and ozone. oxides in coal fly ash on photocatalytic activity. J. Powder Technol. 146, 154–159.
J. Water Resour. 38, 414–422. Yu, X., Liu, S., Yu, J., 2011. Superparamagnetic γ-Fe2O3@SiO2@TiO2 composite
Wang, W.Y., Ku, Y., 2007. Effect of solution pH on the adsorption and photocatalytic microspheres with superior photocatalytic properties. J. Appl. Catal. B–Environ.
reaction behaviors of dyes using TiO2 and Nafion-coated TiO2. Colloids Surf. J. A 104, 12–20.
Physicochem. Eng. Asp 302, 261–268. Zarei, M., Khataee, A.R., Ordikhani-Seyedlar, R., Fathinia, M., 2010. Photoelectro-Fenton
Wang, K.H., Hsieh, Y.H., Chao, P.W., Cgang, C.Y., 2002. The photocatalytic degradation combined with Photocatalytic process for degradation of an azo dye using supported
of trichloroethane by chemical vapor deposition method prepared titanium dioxide TiO2 nanoparticles and carbon nanotube cathode: neural network modeling.
catalyst. J. Hazard Mater. 95, 161–174. J. Electrochemical Acta 55, 7259–7265.
Wang, M., Sun, L., Cai, J., Huang, P., Su, Y., Lin, C., 2013. A facile hydrothermal Zhang, T., Tu, Z., Lu, G., Duan, X., Yi, X., Guo, C., Dang, Z., 2018. Removal of heavy
deposition of ZnFe2O4 nanoparticles on TiO2 nanotube arrays for enhanced visible metals from acid mine drainage using chicken eggshells in column mode. J. Environ.
light photocatalytic activity. J. Mater. Chem. 1, 12082–12087. Manag. 188, 1–8.
Wang, F., Li, M., Yu, L., Sun, F., Wang, Z., Zhang, L., Zeng, H., Xu, X., 2017. Corn-like, Zhao, Y., Liu, F., Qin, X., 2017. Adsorption of diclofenac onto goethite: adsorption
recoverable γ-Fe2O3@SiO2@TiO2 photocatalyst induced by magnetic dipole kinetics and effects of PH. J. Chemosphere 180, 373–378.
interactions. J. Sci. Rep. 7, 1–10. https://doi.org/10.1038/s41598-017-07417-z. Zhao, Z., Long, Y., Luo, S., Wu, W., Ma, J., 2019. Preparation of a magnetic mesoporous
Xu, S.H., Shangguan, W.F., Yuan, J., Chen, M.X., Shi, J.W., 2007. Synthesis and Fe 3 O 4–Pd@ TiO 2 photocatalyst for the efficient selective reduction of aromatic
performance of novel magnetically separable nanospheres of titanium dioxide cyanides. New J. Chem. 43, 6294–6302.
photocatalyst with egg-like structure. J. Appl. Catal. B 71, 177–184. Zuorro, A., Di Battista, A., Lavecchia, R., 2013. Magnetically modified coffee silverskin
Yang, S.T., Zhang, W., Xie, J., Liao, R., Zhang, X., Yu, B., Wu, R., Liu, X., Li, H., Guo, Z., for the removal of xenobiotics from wastewater. J. Chem. Eng. Trans. 35,
2015. Fe3O4@SiO2 nanoparticles as a high-performance Fenton-like catalyst in a 1375–1380. https://doi.org/10.3303/CET1335229.
neutral environment. J. RSC Adv. 5, 5458–5463.

11

Vous aimerez peut-être aussi