You are on page 1of 22

Applied Catalysis A: General 510 (2016) 134–155

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Heterogeneous photocatalytic studies of analgesic and non-steroidal


anti-inflammatory drugs
Amandeep Kaur a , Ahmad Umar b,c,∗ , Sushil K. Kansal a,∗∗
a
Dr S.S. Bhatnagar University Institute of Chemical Engineering & Technology, Panjab University, Chandigarh 160014, India
b
Department of Chemistry, College of Science and Arts, Najran University, PO Box-1988, Najran 11001, Saudi Arabia
c
Promising Centre for Sensors and Electronic Devices (PCSED), Najran University, PO Box-1988, Najran-11001, Saudi Arabia

a r t i c l e i n f o a b s t r a c t

Article history: The presence of pharmaceutical compounds in natural water is considered as an environmental problem
Received 23 July 2015 due to their regular input in aquatic system and hence their removal becomes a challenging task. The
Received in revised form 21 October 2015 present study provides information about the occurrence of pharmaceutical contaminants in water and
Accepted 6 November 2015
their treatment. The advancement of distinct treatment processes for the removal of active pharmaceu-
Available online 6 December 2015
tical ingredients in wastewater has been described. The main objective of this study is to summarize the
latest studies of the degradation of analgesic and anti-inflammatory drugs by heterogeneous photocatal-
Keywords:
ysis. This study also highlights the effect of various operational parameters on degradation efficiency,
Analgesic
Anti-inflammatory drugs
kinetics and mineralization of various pharmaceutical compounds during the process.
Advanced oxidation processes © 2015 Elsevier B.V. All rights reserved.
Photocatalysis
Titania

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
1.1. Occurrence of pharmaceuticals in the aquatic environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
2. Pharmaceutical wastewater treatment processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.1. Conventional treatment processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.2. Advanced oxidation processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
3. Treatment of analgesic and non-steroidal anti inflammatory drugs by heterogeneous photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.1. Ibuprofen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.2. Diclofenac . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.3. Paracetamol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
3.4. Naproxen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
3.5. Other drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

1. Introduction

Water, a natural resource is essential for the existence of life.


∗ Corresponding author at: Department of Chemistry, College of Science and Arts, Although water is present in abundance, but only a limited amount
Najran University, PO Box-1988, Najran 11001, Saudi Arabia. of usable water is available. The level of ground water is decreasing
∗∗ Corresponding author at: Dr. S.S. Bhatnagar University Institute of Chem-
day by day because of more industrialization, improved agriculture,
ical Engineering & Technology, Panjab University, Chandigarh 160014, India.
Fax: +91 172 2534920.
inappropriate disposal techniques, excessive population growth
E-mail addresses: ahmadumar786@gmail.com (A. Umar), sushilkk1@pu.ac.in, and change in climatic conditions [1]. Therefore, the optimum uti-
sushilkk1@yahoo.co.in (S.K. Kansal). lization of the limited water available becomes utmost priority

http://dx.doi.org/10.1016/j.apcata.2015.11.008
0926-860X/© 2015 Elsevier B.V. All rights reserved.
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 135

and this could be done only by improving the quality of water ceutical contaminants. In general, all AOPs possess similar chemical
and removing unwanted organic, inorganic and mineral substances property, i.e. generation of reactive oxygen species which are non
from water streams. selective in nature and have an ability to oxidize recalcitrant com-
Pharmaceuticals have been identified as potential contami- pounds by hydroxylation or dehydrogenation which are further
nants in aquatic environments among various water pollutants mineralized into end products as CO2 and H2 O [25–30]. The treat-
throughout the world [2–5], and pollution by these drugs increased ment of pharmaceutical wastewater by AOPs such as photolysis,
noticeably with their incessant use over the years. The consump- photocatalysis, ozonation, ultrasound radiation, fenton and photo-
tion rate of the drugs has been increasing in the USA, Japan, fenton, etc. exhibit fluctuating degree of removal efficiency [31–38].
France, Spain, United Kingdom and Italy by 11.9% every year. Among all AOPs, heterogeneous photocatalysis is the most distinc-
There are around 3000 compounds used as pharmaceuticals; anti- tive technique for the removal of persistent contaminants from
inflammatory/analgesic and antibiotics being the most common wastewater. Several toxic organic or inorganic water pollutants can
drugs consumed for therapeutic use [6]. be completely mineralized into more biologically degradable and
In the 1960s, existence of pharmaceutical contaminants was harmless substances at comparatively lower costs using heteroge-
detected in wastewater in Europe and United States for the first neous photocatalysis [39].
time. Since then contamination of the environment with these Recent studies demonstrate that non-steroidal anti-
pharmaceuticals has been acknowledged worldwide [7]. The phar- inflammatory drugs (NSAIDs) are frequently found in surface
maceutical industry uses active pharmaceutical ingredients (APIs) water, sewer water, ground water and even in drinking water
for the production of desired drugs, which are pharmacologically as pharmaceutically active compounds [21,40–42]. The NSAIDs
active, toxic in nature [8], resistant to degradation, chemically per- consists of a heterogeneous family of drugs having analgesic
sistent in aqueous media, and has adverse happenings in water and anti-inflammatory features. Among NSAIDs, diclofenac and
bodies and adverse impact on living beings. These compounds ibuprofen are the most commonly consumed drugs and are usually
enter into natural water streams from a number of sources such as found in surface water. The main aim of this review is to assess the
pharmaceutical industrial discharge, personal care products, hos- acquirable information on the occurrence and the removal of anal-
pital effluents, therapeutic drugs, sewage systems and excretion gesic and NSAIDs from wastewater through photocatalysis process.
from livestock etc. [9]. They remain in the environment, enter
into the food chain, bioaccumulate, biomagnify and have danger- 1.1. Occurrence of pharmaceuticals in the aquatic environment
ous effects on humans and wildlife. Few adverse health effects of
these pharmaceutical contaminants are genotoxicity, resistance in The production of pharmaceutical compounds is grabbing con-
microorganisms, toxicity in aquatic environments and hormonal siderable attention for promoting better health care facilities
disruption [10,11]. The existence of these drugs, even in low con- worldwide and longer life expectancy of people. As a result, the
centration in water for a prolonged duration is very dangerous and consumption of various pharmaceutical compounds has increased
their removal is a challenging task [12–15]. In addition, there are drastically over the years and has led to generation of large volumes
no legally permissible limits for these pharmaceuticals in waste of waste water whose treatment is a problem in itself. Thou-
water as the exact impact of chronic exposure of pharmaceuti- sand tons of pharmaceuticals are readily taken up by humans and
cals on aquatic environment still remains unknown [7]. It has been animals which are released into the municipal treatment plants
estimated that about 50 percent of the pharmaceutical wastewater through metabolic excretion and various inappropriate disposal
produced throughout the world is released into the environment techniques. Besides, solid waste excreted by animals which is used
without any proper treatment [16]. These pharmaceuticals may as manure for agriculture enters the food chain through crops. The
also be released into the environment by excretion of partially unused and expired drugs from household, hospitals, industries
transformed active pharmaceutical ingredients by human beings and others sources are also continuously released into the envi-
and animals [7,9,17]. ronment. Though, a small proportion of these drugs are reduced
Treatment of pharmaceutical wastewater is necessary because through incineration, but still major quantity of these drugs have
it not only stabilizes the environment and protects the living been discarded to waste disposal area or flushed down via toi-
beings from various water borne diseases, but also makes water let into municipal sewer [43,44]. As these drugs are partially or
available for use in numerous applications such as agriculture irri- impartially soluble in water, they pass through the water treat-
gation, washing purposes and non-potable municipal use [18,19]. ment plants without being eliminated completely. Moreover, the
Most of the pharmaceutical compounds present in wastewater are pharmaceuticals which make their way through the soil via differ-
polar micropollutants which are not easily removed by wastewater ent means remain intact as they are unabsorbed by the soil because
treatment plants (WWTPs) [20]. Consequently, these micropollu- of their polar nature and hence possess strong potential to get into
tants pass through the surface water and after leaching they enter the environment [45–48].
into the ground water. The WWTPs which receive the pharma- Another source of pharmaceutical pollution is the effluent
ceutical polluted water from sewage plants are not capable of from manufacturing units which is treated in effluent treatment
removing these micropollutants because they are designed only plant (ETP). ETP is an assembly of various treatments which are
for the removal of oxygen demand, suspended solids, pathogens, arranged in such a manner that the final product of one stage is
and nutrients from wastewater. There are many studies which used as input of the next stage for remediation. The various stages
clearly mention that conventional treatments such as adsorption, include preliminary, primary, secondary and tertiary treatment.
coagulation, ion flotation and sedimentation used in WWTPs have In addition, disinfectant (chlorine) is added after the tertiary
limited degradability of active pharmaceutical compounds and treatment to eliminate the pathogens which are responsible for
therefore their end product requires further treatment [15,21]. Bio- various diseases. After disinfection, this partially treated water
logical treatment methods have been traditionally used for the containing remnants of pharmaceutical contaminants enters into
management of pharmaceutical wastewater [22]. These treatment surface and ground water which leads to their potential existence
procedures are limited only to effluents which mainly consist of in drinking water as well. Fig. 1 shows possible sources and
biodegradable compounds and are not hazardous to the biological pathways of pharmaceutical compounds entering into the aquatic
culture [23,24]. environment. The presence of these pharmaceutical compounds
Current literature shows that the advanced oxidation processes and their metabolites even in minute quantities in water are a risk
(AOPs) are far more effective and efficient in removal of pharma- to aquatic and human life [49]. The unfavorable effect on aquatic
136 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 1. Possible sources and pathway of pharmaceuticals compounds into the aquatic environment.

environment includes feminization of male fish, renal failure and operational needs as compared to more advanced methods. It is
impairment of gill and lever in fish. more eco-friendly than chlorination but has some disadvantages as
well, like high energy consumption, difficult to operate and gener-
2. Pharmaceutical wastewater treatment processes ates huge amount of sludge [54]. The pharmaceutical contaminants
remain ineffective in AS process under standard operating con-
2.1. Conventional treatment processes ditions [55]. However, higher concentrations of pharmaceuticals
force it to be restrictive. AS treatment is inefficient where COD con-
The conventional treatment processes consist of physical, chem- tent in wastewater is more than 400 mg/L [22]. These biological
ical and biological methods which include various procedures methods do not completely remove the high concentration of pol-
such as preliminary, primary, secondary and tertiary treatment lutants present in wastewater. The removal efficiency of diclofenac
[50]. Primary treatment partially eliminates pharmaceutical con- varied from 0 to 75% with AS treatment method, while that for car-
taminants from wastewater [21] which require further advanced bamazepine was usually below 10% shown in Fig. 2 [56,57]. The
treatment. The dissolved organic compounds in water can be oxi- primary sedimentation technique eliminated ibuprofen from 12%
dized by biological treatment method. The biological treatment to 45%. The concentration after primary treatment decreased from
process involves the use of microorganisms which consume a vari- 20 to 43% with biological treatment technique [58]. Stumpf et al.
ety of organic compounds [51]. Activated sludge (AS) process has [59] investigated the higher removal efficiency of diclofenac (75%),
been by far the most preferred biological method for degradation ibuprofen (75%), ketoprofen (69%), naproxen (78%) and clofibric
of pharmaceutical wastewater due to its large hydraulic retention acid (34%) with activated sludge process.
time [52]. Temperature, dissolved oxygen, pH, hydraulic reten- The specific pharmaceutical compounds which are retained
tion time, organic load, microorganisms and existence of harmful in the wastewater after secondary treatment are eliminated by
substances are the various factors that affect the efficiency of AS tertiary treatment methods such as coagulation and filtration,
process [53]. AS process has a lower installation cost and lesser
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 137

Fig. 2. Removal efficiency distributions of carbamazepine and diclofenac in WWTPs.

activated carbon, membrane separation, chemical removal and Table 1


Classification of advanced oxidation processes.
electrodialysis. However, the efficiency of these methods fluctuates
critically. Activated carbon is another tertiary treatment method Advanced oxidation processes
which is used for removal of natural and synthetic organic species
Homogeneous process Heterogeneous process
from wastewater. It is reported that activated carbon has high ten-
With or without light source • Photocatalytic ozonation
dency to adsorb pharmaceutical contaminants from wastewater
• Heterogeneous photocatalysis
[60–62]. But, the separation of carbon from water is a difficult task. • Photolysis
These methods have various operational difficulties in addition • H2 O2 photolysis
to high capital costs. Thus, these methods may not be appropri- • Fenton and Photo fenton
ate for the complete removal of pollutants from the environment. • O3 /UV
• O3 /H2 O2 /UV
Therefore, there is a significant need to develop a wastewater treat- • O3 /H2 O2
ment process which can remove the pollutants effectively by simple • Catalytic ozonation
method.

2.2. Advanced oxidation processes eralized products [13,64]. The highlight of this process is that it can
propose various approaches for the production of hydroxyl radicals,
As we know, conventional treatment methods are incapable hence enabling suitable agreement with the typical requirements
enough to treat various toxic and non biodegradable pollutants and of treatment methods AOPs can be used as both pretreatment as
in the last couple of decades, a lot of researchers have prominently well as post-treatment option in biological processes [65].
adopted advanced oxidation processes for the degradation of these Advanced oxidation processes can be classified on the basis of
organic pollutants which are difficult to treat. This clearly signi- phase systems, i.e., homogeneous and heterogeneous processes,
fies the importance of AOPs in removing recalcitrant pollutants both of which can be performed with or without the use of light
from wastewater under ambient conditions [63]. The production source [66]. These processes make use of different light source, oxi-
of hydroxyl radicals (• OH) or other species of same reactivity is the dants and catalysts to produce highly reactive hydroxyl radicals for
characteristic feature of most AOPs. This highly reactive hydroxyl the treatment of wastewater. A list of various possible types of AOPs
radical react with all kinds of natural and synthetic organic pollu- is given in Table 1.
tants non-selectively and rapidly, thereby oxidizing the recalcitrant Photolysis involves the breakdown of a target molecule by
organic matter and eventually transforming them into their min- the absorption of light and leads to its photolytic decomposi-
138 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

tion to intermediate products whose degradation ultimately gives were identified during fenton oxidation and further converted into
mineralized end products [67,68]. However, limited studies have mineralized products such as NH4 + , CO2 and H2 O.
been reported to understand the behaviour of pharmaceutical Ozone (O3 ), a very selective oxidant, has gained noteworthy
substances and their photochemistry in aquatic environments is attention in remediation and purification of wastewater [87]. It
largely undiscovered [33,69,70]. Many pharmaceutical compounds removes the micropollutants as well as reduces its color, foam and
are very sensitive to photolytic transformation [71]. The perfor- odour [88]. In water, it either decomposes itself to produce very
mance of the photolytic process can be improved by combining reactive hydroxyl radicals, which brings out the indirect oxidation
light source with hydrogen peroxide (H2 O2 ), followed by photolytic of organic pollutants or attacks specifically on particular functional
dissociation which produces hydroxyl radicals, thereby facilitating groups of organic moieties via an electrophilic mechanism [89].
oxidation process. The important role of hydrogen peroxide is to This process has been implemented to improve the biodegradabil-
promote the photolysis by using UV light ( < 300 nm) as shown ity and performance of the treatment method [90]. In addition, the
in recent studies [33,71,72]. The photolysis with H2 O2 is gener- efficiency of ozone process is increased by combining ozone with
ally accomplished with high intensity mercury vapor lamp with hydrogen peroxide, UV irradiations and catalyst [13]. The limita-
low and medium pressure so as to reduce the required amount of tion of this process is that it enhances the energy requirement by
H2 O2 required. The advantage of this process can be ascribed to 40–50% in comparison to the conventional treatment process.
the fact that H2 O2 , effective source of • OH, is completely soluble Over the last ten decades, a number of studies have been
in water and requires no separation after treatment [73,74]. Alalm reported on the degradation of pharmaceutical wastewater using
et al. [75] investigated the complete degradation of diclofenac and ozone [91,92]. Aguinaco et al. [93] studied the complete degra-
paracetamol achieved at acidic pH (pH 3) in 120 and 60 min of dation of diclofenac (10−4 M) using 1.5 g/L of TiO2 , 10 mg/L O3
irradiation respectively. However, at higher pH the degradation concentration with flow rate of 30 L/h in 6 min and 65 to 75% TOC
efficiency of these drugs decreased remarkably. This could be due removal was obtained after 60 min of reaction. About 32% mineral-
to poor rate of oxidation of H2 O2 at pH > 4, which leads to lesser ization of diclofenac was reached in 90 min with single ozonation
generation of • OH. The organic content present in the water may [94]. Another study investigated the kinetic parameter studies with
also reduce the radical production, thereby decreasing degradation. O3 and O3 /H2 O2 for the removal of pharmaceutical compounds
Another study compared the degradation efficiency of ibuprofen from water [95]. The degradation of drugs like ibuprofen, clofib-
by different photochemical processes. The results demonstrated ric acid and diclofenac has been studied with O3 /H2 O2 bringing
that UV-C irradiation facilitates only 53% of TOC (total organic car- out almost 98% of mineralization [96]. Several sulfonamides and
bon) removal, while H2 O2 photolysis exhibited an increase in TOC trimethoprim were removed (>90%) from river water by ozonation
removal upto 90% for ibuprofen. It is fascinating to study that even (O3 dose = 7.1 mg/L) in 1.3 min [97]. The required quantity of ozone
though the mineralization of ibuprofen obtained by UV-C photol- depends on various parameters such as alkalinity, pH of wastew-
ysis is insufficient for wastewater treatment [76], this TOC value ater, properties of dissolved organic matter and desired removal
is really higher than that achieved by Mendez-Arriaga et al. [77] performance of target compounds [98]. pH plays an important role
under UV-A irradiation. in ozonation process, as both the absorption and reaction rates of
Fenton process is another type of AOPs used for the treatment ozone get influenced by lowering the pH [99]. This process has few
of wastewater [78]. In this process, fenton’s reagent oxidizes in the disadvantages as well, which restricts its use worldwide as it is
presence of ferrous salt and H2 O2 in a simple redox reaction. This is an energy demanding process and is expensive to implement [88].
a metal catalyzed based oxidation reaction in which iron (Fe2+ ) act- Also, the target compounds which are not fully treated can be sim-
ing as a catalyst gets oxidized to Fe3+ . The efficiency of this process ply transformed into another phase resulting in the production of
can be significantly increased by increasing the number of hydroxyl more other harmful substances [92]. The efficiency of this process
ions produced in the presence of UV irradiations and hence so called can be enhanced by combining it with H2 O2 and UV light, which
“photo-Fenton” process [79]. Its efficiency depends on the various is expected to stimulate the elimination rate considerably by gen-
factors such as reaction time, pH of the solution (ranging between 2 erating more hydroxyl radicals in the treatment procedure [65].
and 4), H2 O2 concentration, nature and quantity of organic contam- Ozone based processes such as O3 /H2 O2, O3 /UV and O3 /UV/H2 O2
inants present in wastewater [80]. The advantage of this process is may produce • OH in different ways. These combined methods are
that the degradation of toxic organic pollutants is possible under considered to be most efficient processes for removal of recalcitrant
UV light as well as solar light [81], Also, it requires simple apparatus organic pollutants.
from bench to large scale at normal temperature and pressure [82]. Recently, catalytic oxidation using ozone has emerged to be
Less toxic and biodegradable intermediates are produced as final very noteworthy and valuable research area [100–103]. The pro-
product which further needs biological post treatment [83]. Rad cess includes both homogeneous and heterogeneous catalytic
et al. [84] studied the degradation of paracetamol in aqueous phase ozonation. The principal of homogeneous catalytic ozonation is
using photoFenton in a binary system. The degradation efficiency stimulation of ozone in the presence of metal ions such as Fe(II),
for 20 mg/L concentration of paracetamol and phenol were found to Mn(II), Ni(II), Co(II), Cd(II), Cu(II), Ag(I), Cr(III), and Zn(II), by
be 85 and 95% with cobalt ferrite (0.2 g/L) and H2 O2 (50 mmol/L) at generating hydroxyl radicals [100,103,104]. Heterogeneous pho-
pH 3.5 respectively. In a study conducted by Loaiza-Ambuludi et al. tocatalytic ozonation employs the use of catalysts such as metal
[76], 96% mineralization of ibuprofen (0.2 mM, pH 3) was achieved oxide and supported metal oxides, over which adsorption of ozone
using Fe3+ (0.25 mM) and H2 O2 (10 mM) after 8 h of UV-C irradia- takes place which further decomposed into surface-obligated rad-
tion. Ravina et al. [85] investigated the removal of diclofenac with icals and hydroxyl radical on the catalyst surface [100,105]. In this
photoFenton oxidation using 400 W low pressure mercury lamp in heterogeneous process, the initial performance of hydroxyl radicals
photoreactor. About 90% TOC removal of diclofenac was achieved is principally influenced by the variation in surface characteristics
within 1 h using Fe2+ (14 mg/L), H2 O2 (340 mg/L) at pH 2.8. de Luna of catalysts, keeping in view that the decomposition of ozone sig-
et al. [86] observed the reaction kinetic for the degradation of parac- nificantly depends on point of zero charge of catalyst, which is also
etamol by varied ferrous iron to H2 O2 ratio using Fenton oxidation. important for the initiation. The point of zero charge and density
It was observed that the diclofenac degradation followed 2nd order of hydroxyl groups present on the surface of the catalyst varies
reaction kinetics. The intermediates such as hydroquinone, ben- indigenously with the modification of the catalyst with metals. pH
zaldehydes, benzoic acids and several non-aromatic organic acids plays a crucial role in the determination of charge properties of
the hydroxyl groups catalyst/water interface [102,106,107]. The
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 139

Fig. 4. General Structure of NSAIDs.

hydroxyl radicals (• OH), while electrons react with oxygen and


reduce it into superoxide radical anions [114]. Hydroxyl radicals
are used as the primary oxidant in the phtotcatalytic oxidation of
organic compounds. It has been displayed that repeated attacks of
organic compounds by • OH leads to complete oxidation of these
pollutants to CO2 and H2 O [115–117].
Various parameters affect the performance of photocatalytic
reactions such as catalyst dose, solution pH, concentration of the
catalyst, light intensity, pH of solution and H2 O2 concentration so
on. The use of renewable energy sources is beneficial for photo-
catalysis by economical point of view. Also, the number of studies
on the degradation of pharmaceutical compounds using solar light
photocatalysis has been reported [67,118,119].

3. Treatment of analgesic and non-steroidal anti


inflammatory drugs by heterogeneous photocatalysis
Fig. 3. Schematic mechanism of photocatalytic reaction for the degradation of drug
using TiO2 photocatalyst. A considerable part of pharmaceutical contaminants in wastew-
ater are comprised of analgesic and NSAIDs, which are extensively
hydroxyl groups, mostly have no surface charge, when the pH of used to relieve pain and reduce inflammation. In 1899, first NSAID,
an aqueous solution is close to the point of zero charge, on the con- aspirin was manufactured by German Bayer Company. Currently,
trary, when the pH is above or beneath the point of zero charge, the more than hundred drugs come under the category of NSAIDs and
oxide surface undergoes protonation and deprotonation. This het- they are familiar to be mostly used all over the world as a pain
erogeneous mechanism is under controversies as different catalysts killer [50]. NSAIDs structurally comprise of acidic moieties which
are used and its pathway is so complicated. The variation is basi- are coupled to a planar, aromatic functional groups (Fig. 4). From the
cally due to the differences in the performance of the catalyst, the mechanistic point of view, they inhibit the activity of cyclooxyge-
characteristics of target compounds and the optimum conditions by nase (COX) and thereby transfer arachidonic acid to prostaglandins
different authors, thereby arising the difficulties in understanding and thromboxane thereby relieving from ongoing pain and fever.
the mechanism of catalytic ozonation. Yang et al. [108] studied the About 30 million people consume NSAIDs every day. In Germany,
TOC removal of diclofenac. Authors observed that TOC removal effi- about 180 tons of ibuprofen, 75 tons of diclofenac and 500 tons
ciency was about 20% and 90% with single ozonation and catalytic of aspirin were used in 2001 [120]. More than 78 tons of aspirin,
(MnOx /Al2 O3 ) ozonation with 1.5 g/L of catalyst concentration and 345 tons of ibuprofen and 86 tons of diclofenac were consumed
30 mg/L of ozone concentration. in England in 2000 [121], And 400 tons of ibuprofen, 37 tons of
The use of heterogeneous photocatalysis in the presence of naproxen, 10 tons of diclofenac and 22 tons of ketoprofen were
semiconductor to degrade recalcitrant compounds is reported used in France in 2004 [122]. Ibuprofen, diclofenac, paracetamol,
primarily in 1980s [109]. After that, this study has gained atten- naproxen, aspirin, mefenamic acid and ketoprofen are the most sig-
tion predominantly among other AOPs. This mechanism is based nificantly used NSAIDs, among these, dicolfenac is highly toxic in
on the stimulation of solid semiconductor under the irradia- nature and its acute effects are examined at concentrations less
tions of suitable wavelength. A large number of catalysts were than 100 mg/L [23].
used in photocatalysis, but TiO2 with anatase phase have some In most of the studies, the attention has been on the estima-
important characteristics, such as high stability, cost effective, tion of optimum conditions for the degradation and mineralization
favorable performance and commercially available. TiO2 absorbs of pharmaceutical compounds. The% degradation efficiency can be
light below 400 nm wavelength which facilitates the use of solar determined using the following formula:
light [110]. TiO2 photocatalysis has become an attractive pro- % efficiency = % conversion = % abatement = extent of degrada-
cess to promote the degradation of contaminants in the aquatic tion (%) = (C0 − C)/C0 × 100
environment since it removes pollutants rapidly and efficiently C0 and C are the initial and final concentrations of drug. The
from water through photocatalytic transformations [63,111–113]. study of operational parameter includes the type of catalyst, solu-
Degussa TiO2 is widely used as photocatalyst in environmental tion pH, catalyst loading, initial drug concentration, wavelength,
applications and it shows high reactivity under UV light during light intensity have also been investigated. Authors also discussed
heterogeneous photocatalysis. the proposed degradation pathway for selected active pharmaceu-
From a mechanistic point of view (Fig. 3), when semiconductor tical ingredients by using mass spectrometry.
(TiO2 , ZnO, CdS, WO3 etc.) is exposed to light, the electrons (e− )
get the valence band (VB) is promoted into the conduction band 3.1. Ibuprofen
(CB) of the photocatalyst, leaving photo-generated holes (h+ ). These
generated electron–hole pairs may either undesirably recombine Ibuprofen (IBP) is one of the most commonly used NSAID and
liberating heat or make their separate ways to the surface of the analgesic drug present in water. About 75–90% of the drug gets
catalyst where they react with species absorbed on the surface of eliminated in the wastewater treatment plants [123]. It is biolog-
the catalyst. When the photocatalytic process takes place in aque- ically active and even its small traces have a negative impact on
ous solution, holes react with water and hydroxide ions to form the environment [124]. It has been extensively used for treatment
140 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

of migraine, fever, muscle ache and tooth ache. Several tons of surface properties of titania after Zr doping. After doping, cata-
ibuprofen is prepared worldwide because of its excessive use in lyst behaves like hydrophobic compound. Thus, polar hydrophilic
the treatment of above mentioned ailments. The most prominent intermediates are generated, which are readily discharged from the
metabolites of IBP, namely carboxyl and/or hydroxyl IBP, have been surface of catalyst into drug solution [128]
found as toxic byproducts [125]. This is one of the most frequently In another study, photodegradation of IBP was evaluated
used drugs, and consequently, there has been gained concern to the together with tetracycline (TC) using different sizes of ZnO
photocatalytic degradation of IBP. nanoparticles which were synthesized using ethanol (ZnOe ) and
Achilleos et al. [126] investigated the photocatalytic degrada- water (ZnOw ) as solvents. TEM images (Fig. 8(a,b)) shows that ZnOe
tion and mineralization of ibuprofen (IBP) and carbamazepine (CBZ) nanoparticles were in slighter agglomerated form and particle size
in aqueous phase by TiO2 under UV-A and simulated solar illu- was found to be 30 nm. Besides, the large (100 nm) particles of ZnOw
mination. TiO2 degussa P25 (mixed phase) exhibited maximum were obtained. The influence of catalyst loading, initial concen-
conversion of IBP and CBZ in 120 min of irradiation among other tration of drug, pH and role of adsorption was investigated. The
six types of TiO2 . About 65% and 61% of IBP was degraded with abatement of TC was larger than IBP with ZnOe as compared to
optimum catalyst dose of degussa P25 TiO2 under both UV-A and ZnOw nanoparticles because of zwitterionic and amphiphilic nature
solar light respectively. About 74% of CBZ was degraded with opti- and higher polarity of TC as compared to IBP. ZnOe smaller particles
mum catalyst loading after 120 min of irradiation. This may be due show an improved performance in the degradation of drugs, due to
to its structure having both phases such as anatase and rutile. This the high specific surface area. The formation of IBP by-product is
behavior may be due to lesser electron-hole recombination rate in more with ZnOe than ZnOw after 240 min of photocatalytic treat-
anatase phase. The study also proposed that this treatment can be ment. These intermediates are benzene derivatives and aliphatic
employed as a post-secondary treatment in WWTPs. acids. It is proposed that transformation products like benzene
The studies have revealed that IBP can be completely elimi- derivates are oxidized to radicals and quinonic structure which
nated during photocatalytic degradation of IBP, (DCF and NPX) with are easily fragmented to aldehydes, ketones, alphatic alcohols and
UV/TiO2 . Maximum conversion of IBP (200 mg/L) has been achieved amines (molecular mass <100) and then easily mineralized into CO2
with a catalyst dose of 1 g/L TiO2 after 240 min of irradiation and and H2 O [129] (Table 2).
it follows the first order rate constant of 9.1 × 10−3 min −1 . Also,
Biological treatment after a photocatalytic process is feasible for 3.2. Diclofenac
the elimination of IBP. This IBP transformation is due to • OH attack
on the methylpropyl phenyl positions and/or propanoic acid moi- Diclofenac (DCF) is a derivative of phenylacetic acid which con-
ety. Different organic acids with m/z (162, 178, and 133) such as sists of phenyl and acetic acid moieties [134]. It is commonly used as
formic, propionic or hydroxypropionic acid or sodium salts have antirheumatic, antiarthric and analgesic and approximately around
been identified by LC–MS during decarboxylation and demethyla- 15% of it is excreted without conversion after human ingestion
tion processes (Fig. 5) [35]. Propionic acid biodegrades initialy to [126]. The presence of diclofenac has been found both in water as
formic acid and acetic acid and then to carbon dioxide and water. well as in sewage effluent which causes different adverse effects
Hence, it is finally degraded to simpler and less toxic compounds. [135,136]. The vulture population declines in Pakistan causing eco-
Wang et al. [127] synthesized Ag-AgBr/TiO2 composite by logical imbalance due to DCF residues in dead animals [137].
one pot method and estimated its photocatalytic activity for the Calza et al. [138] studied the degradation of DCF under simulated
degradation of ibuprofen (IBP) under visible light. Almost 98%, solarlight using TiO2 as catalyst, intermediate and mechanistic
80%, 97% and 62% degradation of Ibuprofen was observed with study was also carried out to access the decomposition of phar-
Ag–AgBr/TiO2 after 2 h illumination with white, blue ( = 465 nm), maceutical drug. Response surface methodology was employed to
green ( = 523 nm), and yellow ( = 589 nm) respectively. The evaluate the optimum values of operational parameters such as
higher removal efficiency of IBP was attained with white and blue catalyst dose and drug concentration for the degradation of target
LED light as compared to yellow and green LEDs as Ag and AgBr compound. Complete mineralization of compound was accom-
were photoactive under white and blue irradiations. This could plished within 2 h of irradiation. There were total 11 photoproducts
be due to efficient charge transfer and separation of photoexcited with four transformation compounds of same m/z ratios at 312 pro-
charge carriers in three component system. Mineralization of IBP duced from hydroxylation process (Fig. 9). Simultaneously, aminic
reached upto 80% under LED irradiations of 6 h (Fig. 6) and it also group was entirely converted into inorganic ions with the for-
decreased the toxicity and aromaticity of the degraded products. mation of nitrate and ammonium ions. Moreover, Vibrio fischeri
A smaller extent of formic acid was identified in IBP solution after was used in assessing the ecotoxicity of photocatalytically treated
360 min of photocatalytic reaction. Further, it was evaluated that solutions. Results demonstrated that toxicity of treated solution
composites were highly effective for Escherichia coli inactivation decreased reached a final value less than 1% showed the efficiency
as compared to conventional Ag–AgBr/P25 and pure titania under of the photocatalytic process in the detoxification of the irradiated
white LED irradiation. solution.
The performance of Zr-doped titania photocatalyst on IBP degra- Achilleos et al. [139] investigated the effect of various param-
dation was observed by changing operational reaction parameters eters such as different types of TiO2 , catalyst loading, H2 O2
such as substrate concentration, catalyst concentration, pH, cata- concentration, initial concentration and water matrix on the degra-
lyst to substrate ratio and catalyst reuse. Conversion of ibuprofen dation of diclofenac (DCF). All experiments were performed in UV-A
increased with increase in catalyst to substrate ratio from 1:1 to 4:1 light affixed immersion well type reactor at pH of 6 with continu-
which was due to diminished effect of catalyst poisoning by using ous oxygen supply. TiO2 degussa P25, Tronox AK-1 and Hombikat
catalyst in large amount. Simultaneously, more intermediates were exhibited 85% conversion of DCF in 4 h of irradiation among other
generated. However, concentration of intermediates decreased as types of TiO2 shown in Fig. 10. This may be due to its structure
a result of higher catalyst dose after 180 min of photocatalytic having both phases such as anatase and rutile. Mixed phase TiO2
reaction. Results confirmed that Zr-doped titania exhibited high (anatase:rutile::75:25) has higher photocatalytic activity than pure
catalytic activity at smaller catalyst dose as compared to pure tita- crystalline phase for degradation of drugs in aqueous phase. Pure
nia. The photocatalytic degradation of IBP increased by a factor of 6 anatase is more active than rutile phase. This behavior may be due
with decreased in pH from 9 to 2 as shown in Fig. 7. This improve- to several reasons (i) the distinct position of the conduction band,
ment in photocatalytic activity is observed due to alteration in (ii) lesser electron-hole recombinations in anatase phase and (iii)
Table 2
Summary of heterogeneous photocatalytic studies on ibuprofen (IBP).

Compound Water matrix Catalyst with optimum Reactor with light source Process parameters Discussion Reference
concentration

Ibuprofen (IBP), diclofenac Milli-Q water TiO2 Solar simulated reactor C[IBP] = 25–200 mg/L, - The maximum photocatalytic degradation of ibuprofen was [35]
(DCF) and C[TiO2] = 1 g/L with Xe-Op lamp (1 Kw, C[TiO2] = 1 g/L, achieved with catalyst dose of 1 g/L
naproxen(NPX) 290–400 nm) - Role of dissolved oxygen on photodegradation of IBP was
investigated

Ibuprofen (IBP) — TiO2 Degussa P25 - Solar pilot plant reactors C[IBP] = 20–200 mg/L, - 4 kJ/L of energy supplied for the complete degradation of IBP [130]
C[TiO2] = 0.1 g/L (CPC-35,CPC-6 and PC- C[TiO2] = 0.1–1 g/L (20 mg/L) with TiO2 dose of 0.1 g/L
1.5) - H2 O2 accelerates the photodegradation process, 50% of IBP
- Solar irradiation (200 mg/L) was eliminated in 120 min with 0.1 g/L of TiO2 ,
- 4KJ/L-5KJ/L solar energy 1028 mg/L of H2 O2 , 4.75 kJ/L
consumption

Ibuprofen (IBP) Milli Q water TiO2 Degussa P25 and Xenon arc lamp, 450 W C[IBP] = 0.02–0.09 mM, - Highest degradation rate achieved by combined UV/TiO2 with [124]
Fe(NO3 )3 .9H2 O C[TiO2] = 1 g/L, ultrasound process and rate was found to be 108.4 × 10 −7 min−1
C[TiO2P25] = 1 g/L C[Fe 3+ ] = 0.05–0.045 mM - The ratio of IBP to Fe3+ was kept at 1:0.5

A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155


C[Fe 3+ ] = 0.045 Mm
Ibuprofen (IBP) and Milli Q water and Aeroxide P25 - Solar simulator, 1000 W C[IBP] = 5–20 mg/L, - Performance of photocatalytic process depends on various factors [126]
Carbamazepine(CBZ) treated domestic - UV-A lamp (9 W, C[TiO2] = 50–800 mg/L, such as catalyst dose, pH and drug concentration were investigated
Wastewater 350-400 nm) pH 3–10, - UV-A light and simulated solar were effective for IBP degradation,
H2 O2 dose = 0.07–1.4Mm mainly in the presence of TiO2 Degussa P25

Ibuprofen (IBP) Deionized water Ag-AgBr/TiO2 - Vis-LED photoreactor C[IBP] = 10 mg/L, - Three component systems (Ag–AgBr/TiO2 ) revealed better [127]
C[Ag-AgBr/TiO2] = 0.5 g/L - white, blue LED C[Ag-AgBr/TiO2] = 0.5 g/L photocatalysis results as compared to one and two component
( = 465 nm), green systems
( = 523 nm), and yellow - Degradation of IBP with Ag–AgBr/TiO2 was about 98% under white
( = 589 nm) LED light

Ibuprofen (IBP) Ultrapure water TiO2 (99% anatase) - UV-C Germicidal lamp (9 C[IBP] = 5 mg/L, - TiO2 /UV-C photocatalytic process was more efficient for IBP [131]
C[TiO2] = 120 mg/L W, 200–280 nm) C[TiO2] = 120 mg/L degradation and mineralization
- UV-A Dark light lamp - Generation of OH radical played important role in IBP degradation
(9 W, 315–400 nm) with first order rate constant of 0.054 min −1

Ibuprofen sodium salt Ultra pure water TiO2 Degussa P25 Four UV–vis solarium lamp C[IBP] = 5–60 mg/L, - Complete degradation of IBP-Na (10 mg/L) was achieved with [132]
(IBP-Na) C[TiO2P25] = 40 mg/L (15 W, 320–400 nm) C[TiO2P25] = 10–40 mg/L, 40 mg/L TiO2 Degussa P25 under 120 min of irradiations
pH: 3-9 - TiO2 degussa P25 was adequate catalyst to degrade the IBP under
analyzed conditions

Ibuprofen Ultra pure water Zr-TiO2 Four UV–vis solarium lamp C[IBP] = 10–60 mg/L, - About 85% ibuprofen (10 mg/L) was degraded with Zr-TiO2 (40 mg/L) [128]
C[TiO2P25] = 40 mg/L (15 W, 320–400 nm) C[Zr-TiO2] = 10–40 mg/L, in 180 min
pH :2–9 - ESMS confirmed the degradation of ibuprofen with Zr-TiO2 was
higher but polymeric compounds are formed partially

Ibuprofen (IBP) and Ultra pure water TiO2 Aeroxide P25 - Immersion well lab scale C[IBP] = 10 mg/L, - Higher drug degradation (90% for DCF, 85% for IBP) was achieved in [133]
Diclofenac (DCF) C[TiO2] = 500 mg/L reactor C[TiO2] = 500 mg/L, 2 h, with sonophotocatalysis as compared sonocatalytic and
- UV-A lamp (9 W, pH: inherent pH sonolytic degradation
350–400 nm) - DCF and IBP decomposition followed pseudo first order kinetic
- Solar simulated xenon model
lamp (1 kW, 290 nm)

Ibuprofen (IBP) and Ultra-pure water ZnO nanoparticles (ZnOw ) - Four UV–vis solarium C[IBP] = C[TC] = 5–60 mg/L, - ZnOe showed better photocatalytic degradation as compared to [129]
tetracycline (TC) and (ZnOe ) lamp (60 W) C[ZnOw] = C[ZnOe] = 10 mg/L, ZnOw
C[ZnOw] = C[ZnOe] = 10 mg/L pH: 3–9 - Intermediates such as benzene and phenolic derivatives were
identified during photocatalytic degradation of IBP by negative
ESI-TOF-MS. These derivatives were oxidized to radials and quinonic
structure and easily converted to CO2 and H2 O

141
142 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 5. Photodegradation pathway of ibuprofen with TiO2 .

Fig. 6. IBP and TOC removal efficiencies under different colour of light irradiations.

higher oxygen adsorption capacity of anatase as a result of higher Another study on degradation of DCF under UV/TiO2 and
density of superfacial hydroxyl groups. Furthermore, pure anatase direct photolysis was investigated by Martinez et al. [140]. In
has more surface area as compared to rutile phase and this can also this, effect of various parameters such as catalyst loading, dis-
lead to an improved activity. The results revealed that abatement solved oxygen and H2 O2 concentration on process kinetics were
as well as mineralization of DCF was effective under UV-A/TiO2 compared. UV light was more effective for degradation of tar-
treatment. get compound than near UV–vis. Complete removal of diclofenac
was achieved under optimal conditions (0.5 g/L of synthesized
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 143

Fig. 7. Effect of pH on abatement of IBP (10 mg/L) with Zr-doped titania (20 mg/L) under UV–vis irradiation.

Fig. 8. TEM images of ZnO nanoparticles precipitated from (a) ethanol (ZnOe) and (b) water (ZnOw ).

Fig. 9. Photocatalytic degradation pathways followed diclofenac (DCF).


144 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 10. Effect of types of TiO2 on DCF degradation.

anatase, 50% O2 (v/v)) under UV light, with apparent rate con-


stant 0.9 min−1 . Oxygen increases the photodegradation process
by reducing the unfavourable electron-hole recombination. Eight
photoproducts were formed during the photolysis of diclofenac
with decarboxylation, photocyclization and dehalogenation. The
byproduct, monohalogenated carbazole was observed by photo-
cyclization pathway and this product is highly phototoxic.
Bagal et al. [141] examined effect of operational parameters
such as inlet pressure of fluid (2-4 bar), pH of drug solution (4–7.5),
TiO2 loading (0.05-0.3 g/L) and H2 O2 concentration (0.05-0.3 g/L)
on the decomposition of DCF, with combinational approach of
hydrodynamic cavitation (HC) and heterogenous photocatalysis. A
slit venturi was used as cavitation device in hydrodynamic reactor.
About 95% DCF degradation and 76% TOC reduction was obtained
with HC/UV/TiO2 /H2 O2 system as compared to UV and UV/TiO2
individually under optimized conditions (TiO2 loading = 0.02 g/L,
H2 O2 loading = 0.02 g/L, inlet fluid pressure = 4 bar, pH 4) and
depicted in Fig. 11 and it followed first order kinetic model for DCF
degradation.
By-products generated during photocatalytic oxidation of DCF
were determined. All products have lower retention time than DCF,
proposed that products were less polar than parent compound.
The chromatograph of degraded products and parent compound
disappeared after 60 min of photocatalytic membrane oxidation.
However, it is possible that smaller molecules were formed during
process, recalcitrant compounds could not be detected and TOC
removal was about 70% under optimized conditions (0.5 g/L TiO2 , Fig. 11. Exent of degradation of diclofenac (DCF) using HC/TiO2 /H2 O2 ,
UV/TiO2 /H2 O2 and combined process of HC/UV/TiO2 /H2 O2 .
2.5 mg/L of DCF concentration at pH 6) [145] (Table 3)

even death. It has been detected as one of the anthropogenic com-


3.3. Paracetamol pound in ground and surface water even after its treatment [148].
Concentration of APAP up to 120 ␮g/L was noticed in ground water
Paracetamol also called as 4-hydroxyacetanilide (ACT) or was reported in study [149]
acetaminophen (APAP), which was initially introduced into medi- There are number of studies carried out on the elimination of
cation as an analgesic/antipyretic by Von Mering in 1893 and has APAP by heterogeneous photocatalysis. Desale et al. [150] opti-
been used as pain reliever for house medication more than 30 years mized the catalyst dose and pH of the APAP solution (50 mg/L),
[147]. It is the most widely used drug after aspirin as a substitute using TiO2 degussa P25 in bath reactor. Complete degradation and
in many countries. Its composition cures more than hundred prob- 72% mineralization of APAP (50 mg/L) was achieved under opti-
lems. But its use in excessive amount may cause liver failure and mized conditions (2 g/L of TiO2 degussa P25, pH 3) in 240 min
Table 3
Summary of heterogeneous photocatalytic studies on diclofenac (DCF).

Compound Water matrix Catalyst with optimum Reactor with light source Process parameters Discussion Reference
concentration

Diclofenac (DCF) Ultra pure water TiO2 - Fluorescence lamp with C[DCF] = 5–80 mg/L, - Experimental data was found to be fit in pseudo first [138]
C[TiO2] = 0.2 g/L black light (125 W, C[TiO2] = 0.2–1.6 g/L, order kinetic model
300-420 nm) pH: 2–9 - Higher concentration of DCF (40 - 80 mg/L) with 1.6 g/L
catalyst loading was able to fit in second order kinetic
model

Diclofenac Deionized water, municipal - TiO2 samples: Degussa Immersion type batch C[DCF] = 5–20 mg/L, - Degussa P25 was more effective catalyst for degradation [129]
waste water treatment P25, Hombicat UV100, photoreactor, Radium C[TiO2] = 50–1600 mg/L, of diclofenac as compared to other catalysts
effluent and ground water Aldrich, Tronox AK-1, UV-A lamp (9 W, H2 O2 dose = 0.07–1.4 mM, - Performance of photocatalytic process depends upon
Tronox TRHP-2, Tronox 350–400 nm) pH: 6 various factors such as catalyst to substrate ratio,
TR reaction time, catalyst and water matrix
- C[TiO2P25] = 50 mg/L

Diclofena Milli pore water - TiO2 Degussa P25 UV light -Medium pressure C[DCF] = 8 mg/L, - The effect of light irradiation, catalyst dose, H2 O2 dose [140]

A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155


- C[TiO2P25] = 0.5 g/L mercury lamp, 366 nm, C[TiO2P25] = 0.1–2 g/L, on photocatalytic degradation of DCF was investigated
NUV-vis light–low pH:6,
pressure mercury lamp, H2 O2 dose = 0–5 mM
254 nm
Diclofenac sodium Distilled water TiO2 -Hydrodynamic cavitation C[DCF] = 20 mg/L, - The radical scavengers study demonstrated that free [141]
C[DCF] = 20 mg/L C[TiO2] = 0.2 g/L (HC) reactor with C[TiO2] = 0.05–0.3 g/L, radical attack was important for the degradation of
photocatalysis H2 O2 dose = 0.05–0.3 g/L, diclofenac
- UV light source, 250 W pH: 4–7.5 - Degradation of diclofenac by photocatalysis was almost
two times higher than HC/TiO2
- Photodegradtion data was found to be fit first order rate
constant

Diclofenac(DCF), Urban waste water, Milli TiO2 - Fluorescence lamp with C[AMX] = 10 mg/L, - Degradation of DCF alone and in mixture with AMX and [142]
Amoxicillin(AMX) and pore water C[TiO2] = 0.8 g/L black light (125 W, C[CBZ] = 5 mg/L, CBZ was examined
Carbamazepine(CBZ) 300–420 nm) C[DCF] = 2.5 mg/L, - Degradation rate was slower in real wastewater as
C[TiO2] = 0.2–0.8 g/L, compared to spiked distilled water
pH: 2–9
Diclofenac Ultra pure water and urban TiO2 Degussa P25 (70% - Borosilicate glass photo CDCF = 10−4 –2.7 × 10−4 M, - Complete degradation of DCF was accomplished within [143]
waste water anatase, 30% rutile) reactor, High pressure Ccat = 0.5–2.5 g/l, 6 min and 60–75%
mercury lamp 313 nm CO3 = 5–30 mg/l, - TOC removal was noticed after 60 min of reaction even
(TQ 718) pH: 7 in deionized and urban waste water

Diclofenac (DCF), Ibuprofen Milli-Q water, river water, -Aeroxide tiatnia P25 - Glass bowl batch reactor 2.50 mg/L of drug mixture. - Various composites were produced from TiO2 [144]
(IBP), Carbamazepine sea water, effulent from - C[TiO2P25] = 500 mg/L - UV-A light, 315–400 nm Equal amount of each drug. nanoparticles and their absorption-photocatalytic
(CBZ), Sulfamethoxazole waste water treatment - UV-B light, 280–360 nm C[Drug] = 0.5 mg/L, studies were carried out
(SMX) and Clofibric acid plant - UV-C light, 220–280 nm C[TiO2] = 500 mg/L - UV-C light source was more effective for the removal of
(CFA). drugs in deionized and river water

Diclofenac Ultra pure water, ground TiO2 -Photocatalytic membrane C[DCF] = 2.5 mg/L, - Degradation and mineralization of diclofenac with TiO2 [145]
C[DCF] = 2 mg/L water and tap water C[TiO2] = 0.5 g/L reactor C[TiO2] = 0.3–0.75 g/L, was about 99.5% and 69% under optimized conditions
-4 black light lamp, 24 W, pH: 6 - Effect of water matrix on photodegradation on
365 nm photodegradation of DCF was estimated

Diclofenac, Indigo Double distilled water TiO2 /Nb2 O5 Philips UV lamp, 36 W, 50 mL of stock of mixture: - Tentative path for the photodegradation of DCF, IC and [146]
carmine(IC), artazine C[TiO2/Nb2O5] = 0.6 g/L 200–400 nm C[DCF] = C[IC] = C[ATR] = 5 mg/L, ATR were investigated
C[DCF] = 5 mg/L C[TiO2] = 0.6 g/L - Results demonstrated that composite not only enhanced
C[IC] = 5 mg/L the degradation rate, mineralization was also
C[ATR] = 5 mg/L accomplished

145
146 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 12. Photocatalytic degradation of APAP (50 mg/L) by TiS, 0.1Cu-TS, 1Cu-TS and 10Cu-TS under visible light. The inset is the plot of ln (C0 /C) versus the irradiation time.

Fig. 13. Photodegradation pathway of APAP.

of irradiation. The rate of photocatalytic degradation of APAP p-nitrophenol and then converted into nitrocatechol and hydro-
decreased with increased in pH from 3 to 11. This is due to repul- quinone. Low molecular weight carboxylic acids were obtained
sion between TiO2 and negatively charged APAP and similar study as the end products. Another study reported that Cu-doped TiO2
of pH effect was conducted by Zhang et al. [151]. Several interme- spheres (CU-TS) exhibited higher photocatalytic degradation as
diates such as hydroquinone, benzoquinone, p-aminophenol and compared to undoped TiO2 . Photocatalytic degradation efficiency
p-nitrophenol were generated during the degradation of APAP at increased with Cu doping (0.1 wt%) but decreased when doping
different time periods. The p-aminophenol was easily oxidized to reached to 1 wt%. The optimum Cu-TS loading (0.1 wt%), yielded
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 147

treatment of wastewater, the problem in its separation because of


its small size stimulates the search of alternative materials. The
complete conversion of APAP was achieved in 5 h by any of these
catalysts. But mineralization of APAP was found to be 43.5%, 39.7%
and 30.8% with degussa P25 TiO2 , TiO2 /SiO2 /Fe3 O4 and TiO2 /Fe3 O4
nanoparticles respectively.
Recent research in the heterogeneous photocatalysis demon-
strated that UV-A/LEDs can be used in the place of conventional
mercury lamps. These lamps are eminent for their high energy cost
and amalgamation of noxious energy [156,157]. Gootostos et al.
[158] synthesized a novel photocatalyst K3 [Fe(CN)6 ]/TiO2 via a sol-
gel method and used it for decomposition of APAP under visible
light using LED lights. The study reported the effect of operational
parameters such as initial concentration of drug solution, catalyst
concentration, solution pH, temperature and number of LED lights
for the decomposition of APAP. The experimental results exhib-
Fig. 14. Photocatalytic degradation efficiency with different scavengers: 20 mmol/L
isopropanol, 0.2 mmol/L sodium oxalate, 0.1 mmol/L TEMPOL (4-hydroxy-2,2,6,6-
ited that much higher degradation efficiency and rate apparent
tetramethylpiperidinyloxy) and N2 in the photodegradation of APAP over were obtained with K3 [Fe(CN)6 ]/TiO2 as compared to commer-
as-synthesized ␤-Bi2 O3 nanospheres. cially available TiO2 degussa P25 catalyst. About 98% and 64%
of APAP was degradaed with K3 [Fe(CN)6 ]/TiO2 under blue LED
( = 450 nm) and green LED ( = 550 nm) respectively. The degra-
the 100% degradation after 3 h of visible light irradiations shown in dation efficiency increased with the increase in number of LED
Fig. 12 [148]. lights because that leads to increase in number of photons that
Degradation mechanism Fig. 13 and parameter optimization of could react with catalyst. It also generates more electron-hole pairs,
APAP under UV/TiO2 was described by Yang et al. [152,153]. In this that increases the photocatalytic degradation. The degradation effi-
study, photocatalytic oxidation of APAP using TiO2 (Degussa P25 ciency of APAP decreased with increase in initial concentration of
TiO2 , Germany) as catalyst, under UVA (365 nm) and UVC (254 nm) APAP and temperature. This can be due to increase the feasibility
light sources were compared. It was found that complete degrada- of recombination of charge carriers and desorption of drug species
tion of APAP (4.0 mM, pH 5.6) was observed with UV-C/TiO2 . The at the surface of catalyst. Also, rate constant increased when intial
mineralization rate was found to be 60% in 300 min (up to 80% in pH of solution was increased from 5.6 to 6.9, further it lowered at
450 min) of reaction. The degradation rate was found to increase a high pH (pH 8.3). Moreover, the degradation efficiency enhanced
from 11 × 10−3 min−1 to 16.5 × 10−3 min−1 with an increase in cat- with catalyst dose up to 1 g/L. The degradation of APAP followed
alyst loading upto optimum dose (0.8 g/L), light intensity, addition pseudo-first order kinetics (Table 4).
of oxygen and by varying the pH. However, after the pH value of 9.5,
the degradation was significantly decreased due to the repulsion
between APAP and negatively charged TiO2 . This study also iden- 3.4. Naproxen
tified an intermediate products including carboxylic acid, aliphatic
compounds through hydroxylation, as well ammonium and nitrate Naproxen (NPX) belongs to 2-arylpropionic acid group which is
were formed. a typical NSAID and extensively used for muscle pain or inflamma-
Xiao et al. [154] prepared highly pure and spherical ␤-Bi2 O3 tory disorders. NPX has been found repeatedly in water bodies [44].
nanospheres of diameter approximately 70 nm by solvothermal Approximately 29 percent of NPX is not eliminated by the sewage
process. A maximum conversion of APAP (10 mg/L) was obtained treatment plant (STP) which consequently enters into the aquatic
with ␤-Bi2 O3 (1 mg/mL) under 180 min of visible irradiation with environment thereby, releasing toxic substances and affects the
a first order rate constant of 0.01387 min−1 . The degradation water adversely [21,160]. Considerable information on degradation
was mainly due to holes which were confirmed by conduct- of naproxen by photocatalysis in comparison to direct photolysis
ing scavenger experiments with sodium oxalate, N2 , TEMPOL is still not available [161]. The photolysis studies have been inves-
(4-hydroxy-2,2,6,6-tetramethylpiperidinyloxy) and isopropanol. tigated under UV light, vaccum ultraviolet (VUV) light and direct
With the addition of sodium oxalate, the photodegradation effi- solar irradiations [162–164]. The effect of oxygen on photodegra-
ciency increased significantly as shown in Fig. 14. Further results dation of NPX and their kinetic study under aerated and deaerated
showed that 89.5% of TOC removal was attained after irradiation for conditions was investigated by Marotta et al. [165]. It was reported
240 min. The apparent reaction rate constant for the degradation of that more toxic derivatives were produced by photodegradation
APAP with prepared ␤-Bi2 O3 was higher as compared to N-doped than the actual compound [166]. This study compared the degrada-
TiO2 , synthetic ␣-Bi2 O3 and Degussa TiO2 P25. Intermediate study tion of NPX by photolysis and photocatalysis under simulated solar
observed the presence of hydroquinone after the photocatalytic light. About 40% removal with 20% mineralization and 90% removal
oxidation of APAP. Then, the aromatic structures of intermediates with 5% mineralization of NPX solution were achieved by pho-
were oxidized into carboxylic acid and CO2 . Ultimately, the APAP tocatalysis and photolysis respectively. Percentage degradation of
was mineralized into small inorganic molecules. NPX by means of photocatalysis decreased due to low adsorption of
Alvarez et al. [155] synthesized magnetic nanoparticles such drug on catalyst surface or reduction of hydroxyl attack. The degra-
as TiO2 /SiO2 /Fe3 O4 and TiO2 /Fe3 O4 by ultrasonic-assisted sol–gel dation rate of NPX increased with TiO2 dose (0.1–1 g/L) for initial
method and used as photocatalyst for the degradation of APAP concentration (0.8 mmol/L) of NPX. The intermediates formed after
mixtures from aqueous phase under UV-A light. TiO2 /SiO2 /Fe3 O4 180 min of photocatalytic reaction were identified by LC/ESI-TOF-
nanoparticles were more stable in aqueous solution under UV light MS. Demethylation and decarboxylation were two prime pathways
as compared to TiO2 /Fe3 O4 nanoparticles. Results demonstrated in recommended for the degradation of NPX. The toxicity of the par-
Fig. 15, that synthesized magnetic nanoparticles showed almost tially decomposed NPX sample was increased that can lead to the
comparable catalytic activity to TiO2 degussa P25. Even though, production of end products which are more toxic than parent NPX
TiO2 degussa P25 was most commonly used photocatalyst for the compound [167].
148
Table 4
Summary of heterogeneous photocatalytic studies on acetaminophen (APAP).

Compound Water matrix Catalyst with optimum Reactor with light source Process parameters Discussion Reference
concentration

Acetaminophen (NSAID) Distilled water TiO2 Photochemical Reactor. C[APAP] = 25–100 ␮M, - About 95% of [151]
C[TiO2] = 1 g/L -Metal Halide lamp C[TiO2] = 0.25–1.0 g/L, paracetamol (100 ␮M)
(250 W, ␭ ≥ 365 nm) pH: 3–11 was degraded with a
TiO2 dose of 1.0 g/L
under 100 min of
irradiation.
- Pseudo first order rate
constant for different
pH:
At pH 3.5, k = 8.09 ␮M min
−1

A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155


pH 6.9, k = 2.84 ␮M min −1
pH 9.5, k = 2.96 ␮M min −1
Acetaminophen Millipore Q water TiO2 Degussa P25(80% Photochemical reactor C[APAP] = 2–10 mM, - In the presence of UV-A [152]
anatase and 20% rutile) (annular cylindrical reactor C[TiO2] = 0.25–1.0 g/L, light, no degradation of
C[TiO2] = 0.4 g/L with a quartz sleeve) pH: 5.1–5.6 acetaminophen was
- Black light blue UVA lamp accomplished without
(8 W, 365 nm) and UVC catalyst
lamp (15 W, 254 nm) - TOC removal was found
to about 60% after
180 min (up to 85% in
450 min) of irradiation
(UV-C) with TiO2
Degussa P25

Acetaminophen Not given ␤-Bi2 O3 Photochemical reactor C[APAP] = 10 mg/L, - Photodegradation of [154]
C[␤-Bi2O3] = 1 g/L - Xenon lamp (1000 W, C[TiO2] = 1 g/L APAP over ␤-Bi2 O3
420 nm cutoff filter) under visible light was
evaluated and compared
with Degussa P25
- Photodegradation data
was found to be fit in
pseudo first order kinetic
model

Acetaminophen Distilled water TiO2 / KAI(SO4 )2 and TiO2 / Lab scale reactor C[APAP] = 0.1–1.0 mM, - Degradation of APAP [159]
NaAIO2 -Blue color LED C[TiO2/KAI(SO4) 2] = 0.5-1.5 g/L, with TiO2 / KAI(SO4 )2
- C[TiO2/KAI(SO4) 2] = 1.0 g/L (440–490 nm), Green color C[TiO2/NaAIO2] = 0.1–1.0 g/L, was higher than other
-C[TiO2/NaAIO2] = 1.0 g/L LED, (510–550 nm) pH: 3–8.1 catalysts such as
Degussa-P25, TiO2 /
NaAIO2 and undoped
TiO2 under blue LED light
- Kinetic study showed
that degradation of APAP
followed first order
kinetic model
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 149

Fig. 15. Photodegradation of acetaminophen in aqueous solution in the absence (blank) and presence of different nanoparticles, Reaction conditions: T ∼ 25◦ C; pH 7;
CAC ,0 = 30 mg/L; catalyst and dose; (䊐) none (blank experiment); (䊏) magnetic iron oxide, 1.0 g/L; (䊉) Degussa P25 TiO2 , 0.5 g/L; () TiO2 /Fe3 O4 , 1.16 g/L; () TiO2 /SiO2 /Fe3 O4 ,
1.34 g/L.

A recent study investigated the photodegradation of naproxen


and three other compounds in water using P25-TiO2 /tetraethyl
orthosilicate (TEOS) film under visible light. The effect of various
parameters such as initial pH of drug solution, catalyst dose, adsorp-
tion, and irradiation time of photocatalytic reaction was studied.
The maximum degradation of naproxen, ibuprofen, diclofenac and
salicyclic acid under optimum conditions (pH 6, 4 pieces of P25-
TiO2 /TEOS films) were 94%, 85%, 65%, and 76% respectively shown
in Fig. 17. This is due to change in surface charge of catalyst (the
pH of TiO2 solution is below 6.5 which is above the pKa values
of NSAIDs). TiO2 surface is positively charged and NSAIDs being
anionic in nature are adsorbed on the surface of catalyst because of
electrostatic attraction and ultimately leading to maximum degra-
dation. It was examined that the degradation of drugs increased
Fig. 16. The schematic representation of formation process of the SiO2@Au@TiO2 with irradiation time and maximum degradation was achieved in
core–shell nanostructures. 10 h of photocatalytic reaction. The photocatalytic degradation of
drugs followed first order kinetics. The intermediate study exhib-
ited that naproxen is photdegraded into nontoxic micromolecules
Couple of studies on photocatalytic degradation of NPX com- with P25-TiO2 /TEOS film [171].
bined with other drugs was reported in literature [168,169]. Ye
et al. [170] studied the degradation of naproxen with prepared 3.5. Other drugs
SiO2 @Au@TiO2 core shell nanostructures under visible light irra-
diations. This experiment was performed in a batch photoreactor Ketoprofen comes under the category of pharmaceutically
and irradiated with xenon lamp (<420 nm). In this study, firstly active compounds, extensively used for the treatment of pains
monodisperse SiO2 cores were synthesized and then decorated during musculoskeletal fractures. Wastewater treatments plants
with Au nanoparticles under ultrasonication. The deposition of Au do not completely eliminate this compound by sorption process
nanoparticles produces the visible light energy by their plasmonic because of its hydrophilic nature; its separation being mainly
effects and intensifies the charge separation. Finally, TiO2 layer was dependent on biological and chemical processes. Feng et al. investi-
coated on the surface of SiO2 @Au collide and after calcination at gated the complete degradation and kinetic study of ketoprofen by
500◦ C for 2 h (shown in Fig. 16), the amorphous TiO2 is converted electro-fenton and anodic oxidation methods [172]. The degrada-
into anatase TiO2. TiO2 coating in nanostructure reduces the loss of tion of ketoprofen followed second order rate constant with a value
Au nanoparticles after use. About 77% photocatalytic degradation of of 2.8(±0.1) × 109 M−1 s−1 . Kim et al., exhibited that the degrada-
naproxen was achieved in 6 h of irradiation with 0.1 wt% Au loaded tion of ketoprofen under UV light (intensity 0.384 mW/cm2 and
SiO2 @Au@TiO2 . This facilitates charge separation by acting as elec- 0.388 mW/cm2 ) was about 90% in 10 min of irradiation [173].
tron reservoir and because of plasmonic effects Au loading leads Aspirin comes under the category of NSAIDs and is used for the
to enhanced spectral response into visible light region. A photo- treatment of fever, pain and arthritis etc. Aspirin is also very effec-
catalytic degradation study on NPX with other three drugs namely tive on blood platelets as it avoids the formation of blood clot inside
ibuprofen, clofibric acid and diclofenac revealed that first order rate the arteries especially in patients suffering from atherosclerosis due
constant for naproxen was higher than others. to its anti platelet effect in the blood. This property of aspirin is very
150 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 17. Photodegradation of four drugs under optimized conditions (pH 6, 4 pieces of P25-TiO2 /TEOS films) in 10 h.

crucial in avoiding heart attack and stokes. About 40,000 tons of this nia was twice than acyclovir and was toxic in nature. Another study
drug is consumed every year and its some residues are getting into reported complete degradation of acyclovir achieved with TiO2 P25
the surface water. in 120 min of irradiation but the inhibition efficiency of treated
Investigation on photocatalytic degradation of aspirin with TiO2 sample was about 10.8%, 7.2% and 14.4% for Selenastrum capricor-
nanoparticles was performed under UV light, the effect of various nutum (96 h), Photobacterium phosphoreum (15 min) and Daphnia
process parameters such as catalyst dose, temperature, pH and time magna (48 h), respectively shown in Fig. 19. However, the toxic
on degradation was studied as well [174]. Optimum conditions for product could be removed with further increase in the irradiation
degradation were determined by using response surface method- duration and toxicity of treated sample decreased considerably as
ology. About 73% of aspirin (5 mg/L, pH 5.5) was decomposed in TOC reduction, but the final degraded solution had small inhibition
50 min with a catalyst dose of 50 mg/L. effect on aquatic life [177].
Recent studies highlighted the effect of cross linked bonding of Yang et al. [178] investigated degradation of three ␤-
TiO2 immobilized polymeric catalyst on aspirin degradation under blockers (metoprolol, propranolol and atenolol) with UV/TiO2 and
both UV and solar light irradiations.TiO2 is able to use sunlight and found that propranolol degraded effectively than metoprolol and
air to produce many reactive species, but filtration of TiO2 slurry atenolol. These three ␤-blockers (100 ␮M, pH 7) were eliminated
from purified water is expensive and time consuming. In this study, completely in 40 min of irradiation using 2.0 g/L of TiO2 and the
photocatalytic degradation of aspirin was carried out in batch reac- degradation kinetics found to be fit the pseudo-first-order kinetic
tor with five types of polymeric/TiO2 catalyst films. These five types model. The degradation of ␤-blockers were mainly due to • OH.
of polymeric cross linked films were prepared by different methods The byproducts were completely mineralized into NH4 + , and NO3
such as freeze dried, heat treated, UV treated, acetaldehyde treated (smaller extent) in 240 min of reaction.
and freeze dried UV treated Fig. 18. Observation showed that freeze Limited research has been dedicated so far for the treatment of
dried cross linked film had a porous structure and high degradation drug ketorolac tromethamine using quantum dots (QDs). Though, it
efficiency in comparison to others [175]. can give very good results. Kaur et al. carried out experiments using
Li et al. [176] studied the degradation of acyclovir with g- TiO2 QDs for degradation of ketorolac tromethamine and found
C3 N4 /TiO2 under visible light. Results exhibited that complete that about 99% of ketorolac tromethamine (10 mg/L, pH 4.4) was
degradation of acyclovir was achieved with g-C3 N4 /TiO2 in 240 min degraded with TiO2 QDs (0.5 g/L) under solar light. The effect of
of irradiation, while very slight degradation was observed with TiO2 operational parameters such as catalyst dose and pH was studied
in 300 min. Authors suggested that, maximum degradation may and shown in Fig. 20. The degradation efficiency increased from
be due to low rate of electron-hole recombination in g-C3 N4 /TiO2 . 83.7% to 97% with increased catalyst dose from 0.25 g/L to 0.5 g/L.
Additionally the generated electron produce reactive species such This was due to increased available active sites on the surface of
as H2 O2 and • O2 . These reactive species decompose acyclovir drug. catalyst. Afterwards, there was decrease in degradation efficiency
However, this degradation process generates three intermediates due to decrease in the penetration of light in the drug suspen-
namely P1 (m/z 214), P2 (m/z 205) and P3 (m/z 152). The photo- sion. The maximum degradation efficiency was achieved at natural
catalytically degraded by-products P1 and P2 were not harmful for pH of drug solution i.e., pH 4.4 The smaller molecules such as K1
three tropic levels such as daphnia, fish and algae as compared (m/z 178), K2 (m/z 228) and K3 (m/z 212) were identified dur-
to parent compound as confirmed by acute and chronic toxicity ing the photocatalytic degradation of parent compound, ketorolac
tests. The aquatic toxicity for intermediate P3 (guanine) on Daph- tromethamine [179].
Table 5
Summary of heterogeneous photocatalytic studies on naproxen (NPX) and other drugs.

Compound Water matrix Catalyst with optimum Reactor with light source Process parameters Discussion Reference
concentration

Equal concentration of 15 Deionied water Anatase TiO2 , rutile TiO2 , Low pressure Hg lamp C[DRUG] = 0.2 mg/L, - More mineralization of [168]
pharmaceutical and TiO2 degussa P25 (264–365 nm) C[TiO2] = 1 mg/L, drugs achieved by TiO2
compounds; Naproxen C[TiO2] = 1 mg/L pH 7.4 photocatalysis as
(NPX) compared to UVC
photolysis
- Antase TiO2 was more
effective to degrade
naproxen, atorvastatin,
ibuprofen, lincomycin,
venlafaxin, diclofenac,
gemfibrozil and
trimethoprim as
compared to rutile TiO2

A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155


Naproxen, malachitegreen, carbamazepineandtheophylline. Water type is not given TiO2 Middle pressure Hg lamp C[NPX] = 15 mg/L, - Photodegradation of [169]
C[TiO2] = 0.2 g/L (100 W, 365 nm) C[TiO2] = 0.2 mg/L, theophylline with TiO2
pH 6.8 under UV light was fast
as compared to naproxen
and carbamazepine
- Effect of initial
concentration of
substrate, pH and
temperature on
degradation was
investigated

Naproxen (NPX) Deionized water SiO2 @Au@TiO2 with Batch photoreactor, Xenon 50 mL of stock of mixture: - SiO2 @Au@TiO2 with [170]
different doping lamp (250 W,  < 420 nm) C[NPX] = 1.0 × 10−5 mol/L, doping concentration of
concentration of Au C[TiO2] = 1.0 g/L, (TiO2 Au (0.1 wt%) having
(0.05–1.0 wt%) dose ∼ 0.30 g/L) better photodegradation
results as compared to
other doping amounts

Naproxen (NPX) and three Water type not given P25-TiO2 /TEOS films Visible light incubator 20 mL of drug solution - The degradations of [171]
other drugs Dose: 1–5 pieces of glazed (10,000 ± 1000 lx) C[NPX] = C[IBP] = C[DCF] naproxen, ibuprofen,
ceramic tiles = C[APAP] = 5 mg/L, diclofenac and salicylic
(2 mm × 2 mm × 5 mm) P25-TiO2 /TEOS films (4 acid, under optimum
pieces of glazed tile), conditions was 94%, 85%,
pH: 6 65%, and 76%
respectively

Ketoprofen (KP) Milli Q water Nanocrystalline TiO2 - Glass immersion C[KP] = 4.8–59 ␮M, - Reaction rate decreased [182]
C[KP] = 59 ␮M C[TiO2] = 0.5 g/L photochemical reactor. C[TiO2] = 0.5–1.0 g/L, with increase in KP
- Medium pressure Hg pH: 7 concentration
lamp, 366 nm. -Low - Photodegradation of KP
pressure Hg lamp, /254 was slower in the
nm presence of oxygen
- Photodegradation of KP
was 1.4 times quicker in
the presence of
synthesized anatase TiO2

151
152 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

Fig. 18. Effect of different cross-linked TiO2 films on the degradation of aspirin under UV light (C0 = 2 mg/L, pH 4.5).

Fig. 19. Growth of acute toxicity assessed with Photobacterium phosphoreum, (b) Selenastrum capricornurum, and (c) Daphnia magna, and (d) TOC reduction and photo-
catalytic degradation of acyclovir.

Chong et al. [180] studied the degradation of carbamazepine in pilot scale and for application to real wastewater treatment. Thus,
synthetic hospital wastewater and suggested that the COD concen- it is concluded that TiO2 based system has a high potential to be uti-
tration in synthetic wastewater is less than real hospital effluent. lized as a sustainable pre-treatment system for hospital wastewater
This study also formed an important basis for future studies based containing drugs and other toxic compounds (Table 5).
on real hospital (mixed) effluents. Choi et al. [181] studied the
degradation of simulated pharmaceutical waste water. They car- 4. Summary
ried out degradation experiments for the simulated waste and real
pharmaceutical waste at similar conditions of pH. The studies car- The occurrence of the pharmaceutical substances in the envi-
ried out by the authors showed that the TiO2 pre-treatment system ronment, specifically in the water bodies has been under the
was capable of removing 78% of carbamazepine, 40% of COD and scanner of the scientific community for over the last couple of
23% of PO4 3− concentrations from the wastewater within 4 h of decades. Pharmaceutical substances in the water bodies repre-
reaction. The proposed treatment poses a combined effect on PO4 3− sent a threat to the ecosystem, due to the fact that they affect
ions reduction and conversion of toxic nitrogen containing com- the human health adversely and their exact impact on aquatic
pounds in wastewater into stable NO3− ions. The studies indicated life still needs to be discovered. Pharmaceutical compounds being
that there is need to study the optimization of parameters for the biologically active substances generally do not biodegrade eas-
degradation of mixture of drugs in water, predominantly at the ily and therefore, are capable of accumulating and persisting in
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 153

Acknowledgements

We greatly acknowledge the financial support obtained under


the UGC Major Research Project Scheme from UGC, Government of
India through a project grant F. No. 41-364/2012(SR) and TEQIP-II
grant of Dr S. S. Bhatnagar UICET, Panjab University, Chandigarh.
Ahmad Umar would like to acknowledge the Ministry of Higher
Education of Saudi Arabia for research fund.

References

[1] S.K. Kansal, A. Kumari, Chem. Rev. 114 (2014) 4993–5010.


[2] F. Sacher, M. Ehmann, S. Gabriel, C. Graf, H.J. Brauch, J. Environ. Monit. 10
(2008) 664–670.
[3] B.J. Richardson, P.K.S. Lam, M. Martin, Mar. Pollut. Bull. 50 (2005) 913–920.
[4] D.W. Kolpin, E.T. Furlong, M.T. Meyer, E.M. Thurman, S.D. Zaugg, L.B. Barber,
H.T. Buxton, Environ. Sci. Technol. 36 (2002) 1202–1211.
[5] H. Yang, G. Li, T. An, Y. Gao, J. Fu, Catal. Today 153 (2010) 200–207.
[6] S.D. Richardson, T.A. Ternes, Anal. Chem. 83 (2011) 4614–4648.
[7] J. Rivera-Utrilla, M. Sanchez-Polo, M.A. Ferro-garcia, G. Prados-Joya, R.
Ocampo-Perez, Chemosphere 93 (2013) 1268–1287.
[8] S. Chelliapan, T. Wilby, P. Sallis, Water Res. 40 (2006) 507–516.
[9] A.M. Deegan, B. Shaik, K. Nolan, K. Urell, M. Oelgemoller, J. Tobin, A.
Morrissey, Int J. Environ. Sci. Technol. 8 (2011) 649–666.
[10] L. Yang, L. Yu, M. Ray, Environ. Sci. Technol. 43 (2009) 460–465.
[11] T. An, H. Yang, W. Song, G. Li, H. Luo, W.J. Cooper, J. Phys. Chem. A. 114
(2010) 2569–2575.
[12] A. Chatzitakis, C. Berberidou, I. Paspaltsis, G. Kyriakou, T. Sklaviadis, I.
Poulios, Water Res. 42 (2008) 386–394.
[13] M. Klavarioti, D. Mantzavinos, D. Kassinos, Environ. Int. 35 (2009) 402–417.
[14] D. Fatta, A. Achilleos, A. Nikolaou, S. Meric, Trends Anal. Chem. 26 (2007)
515–533.
[15] J. Radjenovic, M. Petrovic, B. Damia, Anal. Bioanal. Chem. 387 (2007)
1365–1377.
[16] J.M. Monteagudo, A. Duran, R. Culebradas, I.S. Martin, A. Carnicer, J. Environ.
Manage. 128 (2013), 210-129.
[17] T. An, H. Yang, G. Li, W. Song, W.J. Cooper, X. Nie, Appl. Catal. B: Environ. 94
(2010) 288–294.
[18] H. Cao, X. Lin, H. Zhan, H. Zhang, J. Lin, Chemosphere 90 (2013) 1514–1519.
[19] L.H.M.L.M. Santos, A.N. Araujo, A. Fachini, A. Pena, C. Delerue-Matos,
M.C.B.S.M. Montenegro, J. Hazard. Mater. 175 (2010) 45–95.
Fig. 20. The effect of (a) catalyst dose of and (b) pH of drug solution on the solar-light [20] W.H. Xu, G. Zhang, S.C. Zou, X.D. Li, Y.C. Liu, Environ. Pollut. 145 (2007)
induced degradation of ketorolac tromethamine (20 mg/L). 672–679.
[21] M. Carballa, F. Omil, J.M. Lema, M. Llompart, C. Garcia-Jares, I. Rodriguez, M.
Gomez, T. Ternes, Water Res. 38 (2004) 2918–2926.
[22] D.S. Suman Raj, Y. Anjaneyulu, Process Boichem. 40 (2005) 165–175.
the aquatic environment. In addition, even low concentration lev- [23] E.S. Elmolla, M. Chaudhuri, Desallination 256 (2010) 43–47.
[24] A. Joss, E. Keller, A. Alder, C. McArdell, T. Ternes, H. Siegrist, Water Res. 39
els of the drug in the range of ng/L and ␮g/L, may have serious (2005) 3139–3152.
adverse effects on the environment. Conventional treatment meth- [25] E. Brillas, I. Sires, M.A. Oturan, Chem. Rev. 109 (2009) 6570–6631.
ods such as flocculation, coagulation, filtration and sedimentation [26] L.C. Almeida, S. Garcia-Segura, N. Bocchi, E. Brillas, Appl. Catal. B. 103 (2011)
21–30.
were not effective enough to treat the pharmaceutical wastewater [27] O.K. Dalrymple, D.H. Yeh, M.A. Trontz, J. Chem. Technol. Biotechnol. 82
and therefore required the need to develop alternative meth- (2007) 121–134.
ods. Adsorption is an alternative method for the degradation of [28] A. Umar, R. Kumar, G. Kumar, H. Algarni, S.H. Kim, J. Alloys Compd. 648
(2015) 46–52.
drugs, but not for all prescribed drugs. The literature reports that [29] S. Sood, A. Umar, A. Kaur, S.K. Mehta, S.K. Kansal, J. Alloys Compd. 650 (2015)
activated carbon has high tendency to adsorb pharmaceutical con- 193–198.
taminants, but the separation of carbon from water is a difficult [30] R. Lamba, A. Umar, S.K. Mehta, W.A. Anderson, S.K. Kansal, J. Mol. Catal. A:
Chem. 408 (2015) 189–201.
task. A critical evaluation on the treatment of non steroidal drugs
[31] R. Andreozzi, M. Canterino, R. Giudice, R. Marotta, G. Pinto, A. Pollio, Water
by heterogeneous photocatalysis highlights in this review. TiO2 Res. 40 (2006) 630–638.
photocatalysis is an efficient method for the removal of NSAIDs [32] V.J. Pereira, H.S. Weinberg, K.G. Linden, P.C. Singer, Environ. Sci. Technol. 41
(2007) 1682–1688.
from wastewater. Studies not only describe the reaction kinetics
[33] R. Andreozzi, M. Raffaele, P. Nicklas, Chemosphere 50 (2003) 1319–1330.
but also address the optimization of the operational parameters [34] N. Vieno, T. Tuhkanen, L. Kronberg, Water Res. 41 (2007) 1001–1012.
such as catalyst loading, pH of solution and initial drug concen- [35] F. Mendez-Arriaga, S. Esplugas, J. Gimenez, Water Res. 42 (2008) 585–594.
tration which improves the process efficiency. There are only few [36] T. An, J. An, H. Yang, G. Li, H. Feng, X. Nie, J. Hazard. Mater. 197 (2011)
229–236.
studies on the identification of intermediates formed during the [37] S. Sood, A. Umar, S.K. Mehta, S.K. Kansal, Ceram. Int. 41 (2015) 3355–3364.
degradation reaction. So, there is need to determine the proposed [38] S.K. Kansal, P. Kundu, S. Sood, R. Lamba, A. Umar, S.K. Mehta, New. J. Chem.
photocatalytic degradation pathway of drugs by using photocatal- 38 (2014) 3220–3226.
[39] S.K. Kansal, R. Lamba, S.K. Mehta, A. Umar, Mater. Lett. 106 (2013) 385–389.
ysis treatment. Limited study on the photocatalytic degradation [40] M.J. Benotti, S.A. Snyder, Ground Water 47 (2009) 499–502.
of pharmaceutical drugs using visible and solar light active pho- [41] O.A.H. Jones, N. Voulvoulis, J.N. Lester, Crit. Rev. Environ. Sci. Technol. 35
tocatalyst has been done so far. The major portion of the freely (2005) 401–427.
[42] G.R. Boyd, H. Reemtsma, D.A. Grimm, S. Mitra, Sci. Total Environ. 311 (2003)
available solar energy comprises of visible light. Therefore, in 135–149.
order to enhance the solar efficiency of wide band gap semi- [43] T.J. Scheytt, P. Mersmann, T. Heberer, J. Contam. Hydrol. 83 (2006) 53–69.
conductors under solar irradiation, it is necessary to modify the [44] A. Ziylan, N.H. Ince, J. Hazard. Mater. 187 (2011)
24–36.
nanomaterial in such a way so that it can facilitate visible light
[45] T. Kosjek, E. Hearth, B. Kompare, Anal. Bioanal. Chem. 387 (2007)
absorption. 1379–1387.
154 A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155

[46] S. Mompelat, B. Le Bot, O. Thomas, Environ. Int. 35 (2009) 803–814. [99] M.C. Dodd, M. Buffle, U. von Gunten, Environ. Sci. Technol. 40 (2006)
[47] D. Fatta, A. Achilleos, A. Nikolaou, S. Meric, Trends Anal. Chem. 26 (2007) 1969–1977.
515–533. [100] B. Kasprzyk-Hordern, M. Ziolek, J. Nawrocki, Appl. Catal. B: Environ. 46
[48] T. Heberer, Toxicol. Lett. 131 (2002) 5–17. (2003) 639–669.
[49] Y. Li, G. Zhu, W.J. Ng, S.K. Tan, Sci. Total Environ. 468,469 (2014) 908–932. [101] C.H. Wu, C.Y. Kuo, C.L. Chang, J. Hazard. Mater. 154 (2008) 748–755.
[50] L. Feng, E.D.V. Hullebusch, M.A. Rodrigo, G. Esposito, M.A. Oturan, Chem. [102] L. Zhao, Z.Z. Sun, J. Ma, H.L. Liu, Environ. Sci. Technol. 43 (2009) 2047–2053.
Eng. J. 228 (2013) 944–964. [103] A. Abd El-Raady, T. Nakajima, P. Kimchhayarasy, Ozone: Sci. Eng. 27 (2005)
[51] F. Mohammad, A. Jamrah, Eng. Sci. 35 (1) (2008) 17–25. 495–498.
[52] F.A. El Gohary, S.I. Abou-Elea, H.I. Aly, Water Sci. Technol. 32 (11) (1995) [104] S. Pirgalioglu, T.A. Ozbelge, Appl. Catal. A. 363 (2009) 157–163.
13–20. [105] F.J. Beltran, J. Rivas, P. Alvarez, R. Montero-de-Espinosa, Ozone: Sci. Eng. 24
[53] T. LaPara, A. Konopka, C. Nakatsu, J. Alleman, J. Ind. Microbiol. Biotechnol. 26 (2002) 227–237.
(2001) 203–209. [106] T. Zhang, C.J. Li, J. Ma, H. Tian, Z.M. Qiang, Appl. Catal. B: Environ. 82 (2008)
[54] D. Sreekanth, D. Sivaramarkrishna, V. Himabindu, Y. Anjaneyulu, Bioresour. 131–137.
Technol. 100 (2009) 2534–2539. [107] L. Zhao, Z.Z. Sun, J. Ma, Environ. Sci. Technol. 43 (2009) 4157–4163.
[55] K. Stamatelatou, V. Vavilin, G. Lyberators, Bioresour. Technol. 88 (2) (2003) [108] L. Yang, C. Hu, Y. Nie, J. Qu, Environ. Sci. Technol. 43 (2009) 2525–2529.
137–142. [109] A.L. Pruden, D.F. Ollis, J. Catal. 82 (1983) 404–417.
[56] Y. Zhang, S.-U. Gei␤en, C. Gal, Chemosphere 73 (2008) 1151–1161. [110] R. Goslich, R. Dillert, D. Bahnemann, Water Sci. Technol. 35 (1997) 137–148.
[57] A. Ziylan, N.H. Ince, J. Hazard. Mater. 187 (2011) 24–36. [111] A. Mills, S. Le Hunte, J. Photochem. Photobiol. A. 108 (1998) 1–35.
[58] A.T. Wuersch, L.F.D. Alencastro, D. Grandjean, J. Tarradellas, Water Res. 39 [112] M.R. Hoffmann, S.T. Martin, W.Y. Choi, D.W. Bahnemann, Chem. Rev. 95
(2005) 1761–1772. (1995) 69–96.
[59] M. Stumpf, T.A. Ternes, R.-D. Wilken, S.V. Rodrigues, W. Baumann, Sci. Total [113] H. Hoon, K. Jae-Hong, Water Res. 43 (2009) 2463–2470.
Environ. 225 (1999) 135–141. [114] P. Kundu, A. Kaur, S.K. Mehta, S.K. Kansal, J. Nanosci. Nanotechnol. 14 (2014)
[60] S.A. Snyder, S. Adham, A.M. Redding, F.S. Cannon, J. DeCarolis, J. 1–5.
Oppenheimer, E.C. Wert, Y. Yoon, Desalination 202 (2007) 156–181. [115] P.K. Dutta, S.O. Pehkonen, V.K. Sharma, A.K. Ray, Environ. Sci. Technol. 39
[61] K.J. Choi, S.G. Kim, S.H. Kim, J. Hazard. Mater. 151 (2008) 38–43. (2005) 1827–1834.
[62] Z. Yu, S. Peldszus, P.M. Huck, Water Res. 43 (2008) 1766–1774. [116] M.A. Valenzuela, P. Bosch, J. Jimenez-Becerrill, O. Quiroz, A.I. Paez, J.
[63] R. Andreozzi, V. Caprio, A. Insola, R. Marotta, Catal. Today 53 (1999) 51–59. Photochem. Photobiol. A. 148 (2002) 177–182.
[64] J.Q. Jiang, Z. Zhou, V.K. Sharma, Microchem. J. 110 (2013) 292–300. [117] H. Zhang, G. Chen, D.W. Bahnemann, J. Mater. Chem. 19 (2009) 5089–5121.
[65] I. Kim, N. Yamashita, H. Tanaka, J. Hazard. Mater. 166 (2009) 1134–1140. [118] V. Augugliaro, E. Garcia-Lopez, V. Loddo, S. Malato-Rodriguez, I. Maldonado,
[66] C.P. Huang, C. Dong, Z. Tang, Waste Manage. 13 (1993) 361–377. G. Marci, Sol. Energy 79 (2005) 402–408.
[67] T.E. Doll, F.H. Frimmel, Chemosphere. 52 (2003) 1757–1769. [119] H.M. Coleman, V. Vimonses, G. Leslie, R. Amal, Water Sci. Technol. 55 (2007)
[68] P. Saritha, C. Aparna, V. Himabindu, Y. Anjaneyulu, J. Hazard. Mater. 149 301–306.
(2007) 609–614. [120] T. Ternes, M. Bonerz, T. Schmidt, J. Chromatogr. A. 938 (2001) 175–185.
[69] P. Bartels Jr., W. von Tümpling, Sci. Total Environ. 374 (2007) 143–155. [121] A. Nikolaou, S. Meric, D. Fatta, Anal. Bioanal. Chem. 387 (2007) 1225–1234.
[70] S. Canonica, L. Meuniera, U. von Guntena, Water Res. 42 (2008) 121–128. [122] Y. Kim, K. Choi, J. Jung, S. Park, P.G. Kim, J. Park, Environ. Int. 33 (2007)
[71] S.K. Khetan, T.J. Collins, Chem. Rev. 107 (2007) 2319–2364. 370–375.
[72] H. Shemer, Y.K. Kunukcu, K.G. Linden, Chemosphere 63 (2006) 269–276. [123] C. Tixier, H.P. Singer, S. Oellers, S.R. Muller, Environ. Sci. Technol. 37 (2003)
[73] P.R. Gogate, A.B. Pandit, Adv. Environ. Res. 8 (2004) 553–597. 1061–1068.
[74] M.I. Litter, Introduction to photochemical advanced oxidation processes for [124] J. Madhavan, F. Grieser, M. Ashokkumar, J. Hazard. Mater. 178 (2010)
water treatment, in: P. Boule, D.W. Bahnemann, P.K.J. Robertson (Eds.), 202–208.
Environmental Photochemistry Part II, Springer Berlin Heidelberg [125] S. Mozia, A.W. Morawski, Catal. Today 193 (2012) 213–220.
Publishing, New York, 2005, pp. 325–366. [126] A. Achilleos, E. Hapeshi, N.P. Xekoukoulotakis, D. Mantzavinos, D.
[75] M.G. Alalm, A. Tawfik, S. Oakawara, J. Environ. Chem. Eng. 3 (2015) 46–51. Fatta-Kassinos, Sep. Sci. Technol. 45 (2010) 1564–1570.
[76] S.L. Ambuludi, M. Panizza, N. Oturan, M.A. Oturan, Catal. Today 224 (2014) [127] X. Wang, Y. Tang, Z. Chen, T.T. Lim, J. Mater. Chem. 22 (2012) 23149–23158.
29–33. [128] J. Choina, G.U. Fischer, H. Flechsig, V.A. Kosslick, N.D. Tuan, N.A. Tuyen, A.S.
[77] F.M. Arriaga, S. Esplugas, J. Gimenez, Water Res. 44 (2010) 589–595. Tuyen, J. Photochem. Photobiol. A. 274 (2014) 108–116.
[78] A.J. Watkinson, E.J. Murbyc, S.D. Costanzo, Water Res. 41 (2007) 4164–4176. [129] J. Choina, A. Bagabas, Ch. Fischer, G.-U. Flechsig, H. Kosslick, A. Alshammari,
[79] L. Perez-Estrada, M. Maldonado, W. Gernjak, A. Aguera, A.R. Fernández-Alba, A. Schulz, Catal. Today 241 (2015) 47–54.
M.M. Ballesteros, S. Malato, Catal. Today 101 (3–4) (2005) 219–226. [130] F. Mendez-Arriage, M.I. Maldonado, J. Gimenez, S. Esplugas, S. Malato, Catal.
[80] W. Li, V. Nanaboina, Q. Zhou, G.V. Korshin, J. Hazard. Mater. 244–245 (2013) Today 144 (2009) 112–116.
698–708. [131] J.C.C. Da Silva, J.A.R. Teodoro, R.J.D.C.F. Afonso, S.F. Aquino, R. Augustia, J.
[81] L. Perez-Estrada, S. Malato, W. Gernjak, A. Aguera, E. Thurman, I. Ferrer, A.R. Mass Spectrom. 49 (2014) 145–153.
Fernández-Alba, Environ. Sci. Technol. 39 (21) (2005) 8300–8306. [132] J. Choina, H. Kosslick, Ch. Fischer, G.-U. Flechsig, L. Frunza, A. Schulz, Appl.
[82] V. Kavitha, K. Palanivelu, Chemosphere 55 (9) (2004) 1235–1243. Catal. B 129 (2013) 589–598.
[83] I. Munoz, J. Peral, J. Ayllon, S. Malato, P. Passarinho, X. Domenech, Water Res. [133] I. Michael, A. Achilleos, D. Lambropoulou, V.O. Torrens, S. Perez, M. Petrovic,
40 (19) (2006) 3533–3540. D. Barcelo, D. Fatta-Kassinos, Appl. Catal. B. 147 (2014) 1015–1027.
[84] L.R. Rad, M. Irani, F. Divsar, H. Pourahmad, M.S. Sayyafan, I. Haririan, J. [134] T. Nogrady, D.F. Weaver, Medicinal Chemistry—A Molecular and
Taiwan Inst. Chem. Eng. 47 (2015) 190–196. Biochemical Approach, 3rd ed., Oxford University Press, 2005, P. 525.
[85] M. Ravina, L. Campanella, J. Kiwi, Water Res. 36 (2002) 3553–3560. [135] J.F. Garcia-Araya, F.J. Beltran, A. Aguinaco, J. Chem. Technol. Biotechnol. 85
[86] M.D.G. de Luna, R.M. Briones, C.-C. Su, M.-C. Lu, Chemosphere 90 (2013) (2010) 798–804.
1444–1448. [136] S. Yan, W. Song, Environ. Sci. 16 (2014) 697.
[87] J. Arana, J. Herrera Melian, J. Dona Rodriguez, O. Gonzalez Diaz, A. Viera, J. [137] L.J. Oaks, M. Gilbert, M.Z. Virani, R.T. Watson, C.U. Meteyer, B.A. Rideout, H.L.
Perez Pena, P. Marrero Sosa, V. Espino Jimenez, Catal. Today 76 (2–4) (2002) Shivaprasad, S. Ahmed, M. Chaudhry, M. Arshad, S. Mahmood, A. Ali, A.A.
279–289. Khan, Nature 427 (2004) 630–633.
[88] T. Larsen, J. Lienert, A. Joss, H. Siegrist, J. Biotechnol. 113 (1–3) (2004) [138] P. Calza, V.A. Sakkas, C. Medana, C. Baiocchi, A. Dimou, E. Peelizzetti, T.
295–304. Albaniis, Appl. Catal. B. 67 (2006) 197–205.
[89] T. Ternes, J. Stuber, N. Herrmann, D. McDowell, A. Ried, M. Kampmann, B. [139] A. Achilleos, E. Hapeshi, N.P. Xekoukoulotakis, D. Mantzavinos, D.
Teiser, Water Res. 37 (2003) 1976–1982. Fatta-Kassinos, Chem. Eng. J. 161 (2010) 53–59.
[90] E. Cokgor, I. Alaton, O. Karahan, S. Dogruel, D. Orhon, J. Hazard. Mater. 116 [140] C. Martınez, M. Canle, M.I. Fernandez, J.A. Santaballa, J. Faria, Appl. Catal. B
(2004) 159–166. 107 (2011) 110–118.
[91] H. Nakada, A. Shinohara, K. Murata, S. Kiri, N. Managaki, N. Sato, H. Takada, [141] M.V. Bagal, P.R. Gogate, Ultrason. Sonochem. 21 (2014) 1035–1043.
Water Res. 41 (2007) 4372–4382. [142] L. Rizzo, S. Meric, M. Guida, D. Kassinos, F. Russo, V. Belgiorno, Water Res. 43
[92] R. Dantes, S. Contreras, C. Sans, S. Esplugas, J. Hazard. Mater. 150 (2008) (2009) 4070–4078.
790–794. [143] A. Aguinaco, F.J. Beltran, J.F. Garcia-Araya, A. Oropesa, Chem. Eng. J. 189–190
[93] A. Aguinaco, F.J. Beltran, J.F. Garcia-Araya, A. Oropesa, Chem. Eng. J. 189–190 (2012) 275–282.
(2012) 275–282. [144] N. Rioja, P. Benguria, F.J. Penas, S. Zorita, Environ. Sci. Pollut. Res. 21 (2014)
[94] D. Vogna, R. Marotta, A. Napolitano, R. Andreozzi, M. d’Ischia, Water Res. 38 11168–11177.
(2004) 414–422. [145] V.C. Sarasidis, K.V. Plakas, S.I. Patsios, A.J. Karabelas, Chem. Eng. J. 239 (2014)
[95] K. Ikehata, N. Jodeiri Naghashkar, M. Gamal EI-Din, Ozone: Sci. Eng. 28 299–311.
(2006) 353–414. [146] F.V. de Andrade, G.M. de Lima, R. Augusti, M.G. Coelho, Y.P.Q. Assis, I.R.M.
[96] C. Zwiener, F.H. Frimmel, Water Res. 34 (2000) 1881–1885. Machado, J, Environ. Chem. Eng. 2 (2014) 2352–2358.
[97] C. Adams, M. Asce, Y. Wang, K. Loftin, M. Meyer, J. Environ. Eng. 128 (2002) [147] E. Moctezuma, E. Leyva, C.A. Aguilar, R.A. Luna, C. Montalvo, J. Hazard.
253–260. Mater. 243 (2012) 130–138.
[98] M. Huber, A. Gobel, A. Joss, N. Hermann, D. Loffler, C. McArdell, A. Ried, A. [148] C.J. Lin, W.T. Yang, Chem. Eng. J. 237 (2014) 131–137.
Siegrist, T. Ternes, U. Von Gunten, Environ. Sci. Technol. 39 (2005) [149] A.B. Caracciolo, E. Topp, P. Grenni, J. Pharm. Biomed. Anal. 106 (2015) 25–36.
4290–4299.
A. Kaur et al. / Applied Catalysis A: General 510 (2016) 134–155 155

[150] A. Desale, S.P. Kamble, M.P. Deosarkar, Int. J. Chem. Phys. Sci. 2 (2013) [166] M. Isidori, M. Lavorgna, A. Nardelli, A. Parrella, L. Previtera, M. Rubino, Sci.
140–148. Total Environ. 348 (1-3) (2005) 93–101.
[151] X. Zhang, F. Wu, X.W. Wu, P. Chen, N. Deng, J. Hazard. Mater. 157 (2008) [167] F. Mendez-Arriaga, J. Gimenez, S. Esplugas, J. Adv. Oxid. Technol. 11 (3)
300–307. (2008) 435–444.
[152] L. Yang, L.E. Yu, M.B. Ray, Water Res. 42 (2008) 3480–3488. [168] A. Hu, X. Zhang, D. Luong, K.D. Oakes, M.R. Servos, R. Liang, S. Kurdi, P. Peng,
[153] L. Yang, L. Yu, M. Ray, Environ. Sci. Technol. 43 (2009) 460–465. Y. Zhou, Waste Biomass Valor. 3 (2012) 443–449.
[154] X. Xiao, R. Hu, C. Liu, C. Xing, C. Qian, X. Zho, J. Nan, L. Wang, Appl. Catal. B: [169] R. Liang, A. Hu, W. Li, Y.N. Zhou, J. Nanopart. Res. 15 (2013) 1–13.
Environ. 140–141 (2013) 433–443. [170] M. Ye, H. Zhou, T. Zhang, Y. Zhang, Y. Shao, Chem. Eng. J. 226 (2013) 209–216.
[155] P.M. Alvarez, J. Jaramillo, F. Lopez-Pinero, P.K. Plucinski, Appl. Catal. B. 100 [171] H. Zhang, P. Zhang, Y. Ji, J. Tian, Z. Du, Chem. Eng. J. 262 (2015) 1108–1115.
(2010) 338–345. [172] L. Feng, N. Oturan, E.D. van Hullebusch, G. Esposito, M.A. Oturan, Environ.
[156] P. Xiong, J.Y. Hu, Sep. Purif. Technol. 91 (2012) 89–95. Sci. Pollut. Res. 21 (2014) 8406–8416.
[157] E. Haggiage, E.E. Coyle, K. Joyc, M. Oelgemoller, Green Chem. 11 (2009) [173] I. Kim, H. Tanaka, Environ. Int. 35 (2009) 793–802.
318–321. [174] S. Ghajar, M.R. Sohrabi, J. Chem. Pharm. Res. 4 (2012) 814–821.
[158] M.J.N. Gotostos, C.C. Su, M.D.G.D. Luna, M.C. Lu, J. Environ. Sci. Health Part A [175] D. Mukherjee, S. Barghi, A.K. Ray, Processes 2 (2014) 12–23.
49 (2014) 892–899. [176] G. Li, X. Nie, Y. Gao, T. An, Appl. Catal. B: Environ. 180 (2016) 726–732.
[159] M.L.P. Dalida, K.M.S. Amer, C.C. Su, M.C. Lu, Environ. Sci. Pollut. Res. 21 [177] T. An, J. An, Y. Gao, G. Li, H. Fang, W. Song, Appl. Catal. B: Environ. 164 (2015)
(2014) 1208–1216. 279–287.
[160] K. Press-Kristensen, A. Ledin, J. Schmidt, M. Henze, Sci. Total Environ. 373 [178] H. Yang, T. An, G. Li, W. Song, W.J. Cooper, H. Luo, X. Guo, J. Hazard. Mater.
(2007) 122–130. 179 (2010) 834–839.
[161] E. Felis, D. Marciocha, J. Surmacz-Gorska, K. Miksch, Water Sci. Technol. 55 [179] A. Kaur, A. Umar, S.K. Kansal, J. Colloid Interface Sci. 459 (2015) 257–263.
(2007) 281–286. [180] M.N. Chong, B. Jin, J. Hazard. Mater. 199–200 (2012) 135–142.
[162] D. Ma, G. Liu, W. Lv, K. Yao, X. Zhang, H. Xiao, Environ. Sci. Pollut. Res. Int. 21 [181] J. Choi, H. Lee, Y. Choi, S. Kim, S. Lee, S. Lee, W. Choi, J. Lee, Appl. Catal. B:
(2014) 7797–7804. Environ. 147 (2014) 8–16.
[163] E. Arany, R.K. Szabo, L. Apati, T. Alapi, I. Iiisz, P. Mazellier, A. Dombi, K. [182] C. Martínez, S. Vilarino, M.I. Fernández, J. Faria, M. Canle, J.A. Santaballaa,
Galda-Schrantz, J. Hazard. Mater. 262 (2013) 151–157. Appl. Catal. B. 142–143 (2013) 633–646.
[164] Y. Chen, G. Liu, K. Yao, W. Lu, Chin. J. Environ. Eng. 7 (2013) 473–476.
[165] R. Marotta, D. Spasiano, I. Di Somma, R. Andreozzi, Water Res. 47 (2013)
373–383.

You might also like