You are on page 1of 11

Water Research 164 (2019) 114920

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Degradation of emerging organic pollutants in wastewater effluents


by electrochemical photocatalysis on nanostructured TiO2 meshes
S. Murgolo a, S. Franz b, H. Arab b, M. Bestetti b, E. Falletta c, G. Mascolo a, *
a
CNR, Istituto di Ricerca Sulle Acque, Via F. De Blasio 5, 70132, Bari, Italy
b
Politecnico di Milano, Department of Chemistry, Materials and Chemical Engineering, G.Natta, Milano, Italy
c
Universita di Milano, Dipartimento di Chimica, Milano, Italy

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

Article history: An immobilized photoactive TiO2 coating grown directly on titanium meshes was successfully exploited
Received 13 March 2019 for the electrochemical photocatalytic degradation of carbamazepine in real secondary wastewater
Received in revised form effluent. The catalyst was prepared by Plasma Electrolytic Oxidation and during the photocatalytic water
10 July 2019
treatment an electrical polarization (bias) was applied to the catalyst. The investigated process was
Accepted 26 July 2019
Available online 28 July 2019
compared with the conventional one employing suspended TiO2 powder (Degussa P25). Results showed
that carbamazepine degradation rate follows the order UV/supported TiO2þbias z UV/TiO2 Degussa
P25 > UV/supported TiO2 > UV. The investigation also included the identification of other micro-
Keywords:
Immobilized catalyst
pollutants and degradation products. This allowed the detection of 201 compounds present in the sec-
Titanium dioxide ondary wastewater effluent employed for the photocatalysis tests, 51 of them also successfully associated
Compounds of emerging concern to compounds of emerging concern (CECs), and 194 to transformation products (TPs). The degradation of
Secondary wastewater effluent detected compounds followed first-order kinetics and the mean kinetic constant values of the 51 CECs
Non-target screening resulted to be 0.048, 0.035 and 0.043 min1 for the TiO2þBias þ UV, TiO2þUV and UV, respectively. As for
Transformation products TPs, results showed that the TiO2þBias þ UV treatment is much more efficient than both TiO2þUV and
UV in minimizing the intensity of the organics in the real wastewater. Such a better performance was
more pronounced at higher reaction time reaching 60% reduction of mean peak area of TPs at 90 min of
reaction. Among the detected TPs also compounds belonging to known carbamazepine TPs were found.
This allowed to propose a degradation pathway of carbamazepine. The supported catalyst was positively
tested for 15 cycles demonstrating that it has the potential to be used in real wastewater tertiary steps
aimed at removing CECs.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction 2019; Stalter et al., 2010; Ternes, 1998; Zwiener et al., 2002).
Wastewater reuse has been prompted as a new challenge due to
Water resources and societal wellbeing are currently threatened the ever increasing demand of fresh water caused by the growth of
by contaminants of emerging concern (CECs) as well as by patho- population and the high consumption in agricultural and industrial
gens including antibiotic resistant bacteria and viruses. The impact sectors. Water recycling for non-potable uses from conventional
of wastewater effluents containing CECs on the quality of receiving water treatment processes is applied in water-scarce regions and it
water bodies has been widely demonstrated (Kasprzyk-Hordern can be among the driving forces for the requirement of water of
et al., 2009; Zhou et al., 2009). Removal of CECs and/or their me- better quality. The relevance of addressing the issue of CECs was
tabolites is related to both the nature of the specific organics and to acknowledged by the Directive 2013/39/EU listing priority sub-
the treatment methods employed in the treatment plant stances and further supported by the implementation of Decision
(Castiglioni et al., 2018; Joss et al., 2005; Loos et al., 2013). CECs can (EU) 2015/495 on March 20, 2015 establishing a watch list including
be also degraded into products that are still active (Rizzo et al., 10 target substances for European Union-wide monitoring
€der et al., 2016; Sousa et al., 2018). The watch list was
(Schro
recently revised by the Decision (EU) 2018/840. It follows that
requirement of water of better quality acts as a driving force for the
* Corresponding author.
application of post-treatment technologies in wastewater
E-mail address: giuseppe.mascolo@ba.irsa.cnr.it (G. Mascolo).

https://doi.org/10.1016/j.watres.2019.114920
0043-1354/© 2019 Elsevier Ltd. All rights reserved.
2 S. Murgolo et al. / Water Research 164 (2019) 114920

treatment plants (WWTPs). Conventional wastewater treatment tested for the removal of CECs by treating a real secondary waste-
processes lead to incomplete removal of CECs (Luo et al., 2014; Tran water. The wastewater was also spiked with carbamazepine, which
et al., 2018) being not designed to perform such organics removal. is one of the main CECs in municipal wastewaters. In order to get
Advanced biological treatments showed better performance than insights about the degradation pathway in the real water matrix, an
conventional ones for selected CECs removal (Balest et al., 2008). It identification of the transformation products (TPs) formed during
is worth noting that, due to the driving force of preserving the the photocatalytic treatment of secondary wastewater effluent was
environmental quality of receiving water bodies, the environ- performed, too. Finally, the supported catalyst was positively tested
mental legislation of countries such as Switzerland was modified in for 15 cycles demonstrating that it has the potential to be used in
order to achieve 80% removal of among 6 CECs from a list of 12 real wastewater tertiary step aimed at removing CECs.
compounds (Giannakis et al., 2015; Logar et al., 2014). Advanced
oxidation processes (AOPs), including photocatalysis, can remove a
large range of organic micropollutants and CECs (Fagan et al., 2016; 2. Materials and methods
Miklos et al., 2018; Pagano et al., 2008), being also possible their
integration with biological processes (Del Moro et al. 2013, 2016). 2.1. Preparation of nanostructured TiO2 meshes
Heterogeneous photocatalysis has become relevant in the last years
since sometimes chemicals are not necessary for the oxidation Titanium dioxide coatings were obtained by PEO of grade I ti-
processes (Andronic et al., 2016; Cates, 2017; Dong et al., 2015). For tanium expanded meshes in 1.5 M H2SO4 (14.7 wt %), at constant
example, the use of titanium dioxide (TiO2) as a photocatalyst is potential of 150 V, for 5 min. Several smaller samples (18 cm2 area)
receiving growing attention for its effectiveness, due to the were also produced in the same conditions for further character-
nonspecific nature of the reactive species produced under UV ization. Visually, all the samples had a light gray and uniform
irradiation (Nakata and Fujishima, 2012; Ochiai and Fujishima, appearance. The electrolyte temperature was set at about 5  C by
2012). However, it has to be considered that the efficiency of the a cryostat (HAAK D10). After PEO, the samples were rinsed with
process still needs to be improved due to both the low quantum water and dried in a stream of air.
efficiency of the process, namely too many photons are not used to Morphology was investigated by scanning electron microscopy
produce hydroxyl radicals, and possible catalyst poisoning espe- (SEM) using a Zeiss EVO 50 instrument. The phase structure and
cially with real wastewater. In addition, the application of sus- texture of the titanium dioxide films were assessed by X-ray
pended TiO2 is very limited at both pilot and full-scale since the diffraction (XRD) using a Philips PW1830 instrument operating in
small size of the catalyst complicates its recovery at the end of the Bragg-Brentano geometry at a potential of 40 kV with a filament
treatment, reducing its potential reuse and compromising the current of 40 mA. Diffractograms were acquired with Cu Ka1 radi-
quality of treated effluent (Byrne et al., 2017). To overcome this ation in the 2q range 20e60 at the scanning rate of 2.5 per min.
drawback, several authors have studied the immobilization of TiO2 The XRD patterns were indexed according to the powder diffraction
particles on different materials (Borges et al., 2015; Comparelli files of titanium (ICDD-PDF 44e1294), anatase (ICDD-PDF 21e1272)
et al., 2004; Murgolo et al., 2017; Petronella et al. 2011, 2013). and rutile phases (ICDD-PDF 21e1276). The mass fraction of anatase
However, despite the huge number of papers available in the was calculated according to Eq. (1) (Spurr and Myers, 1957), where
literature, real applications of photocatalysis with supported cata- IR is the intensity of the strongest rutile reflection, (110), and IA is
lysts are still very rare. the intensity of the strongest anatase reflection, (101),
Electrochemical TiO2 photocatalysis is a well-known technique
1
among AOPs for water treatment but a not often applied process. fA ¼  % (1)
The photoactive coating is grown directly on the titanium meshes, ðIR Þ
1 þ 1:26 ðI AÞ
leading to good mechanical adhesion to the substrate and good
electrical conductivity. During the wastewater treatment step
Surface porosity was evaluated by analyzing several SEM surface
aimed at removing organic pollutants, an electrical bias can be
micrographs taken at 20,000x magnification using ImageJ analysis
effectively applied to the catalyst through the mesh, leading to a
software. Film thickness and depth profiling were assessed by Glow
synergistic effect with UV light and a faster degradation kinetics
Discharge Optical Emission Spectrometry (GD-OES) using a Spec-
with respect to either photocatalytic or the electrochemical
truma GDA750 analyzer operated at 700 V in argon atmosphere at
process.
230 Pa. The area of analysis was of ~2.5 mm diameter, and the
Photoactive TiO2 coatings are obtained by a number of tech-
monitored light emissions during the analysis were 130 nm and
niques including sol-gel, CVD, RF magnetron sputtering, plasma
362 nm for oxygen and titanium, respectively. The photocurrent
spray, electron beam evaporation, anodic oxidation (AO) (Bestetti
density of the TiO2 film was measured by Linear Sweep Photo-
et al., 2007) and plasma electrolytic oxidation (PEO) (Franz et al.,
Voltammetry (LSPhV) in 4.2 mM KCl aqueous solution with and
2016). Catalysts obtained by PEO and AO were already proved to
without irradiation at a scan rate of 5 mV s1 and room tempera-
be effective in the degradation of a model dye (Franz et al., 2015).
ture using a potentiostat/galvanostat (Solartron Analytical Modu-
The experimental set-up for PEO and AO are similar. However, due
Lab ECS). The reference electrode was a Saturated Calomel
to the high operating voltage applied during PEO, very peculiar
Electrode (SCE). The counter electrode was a Tie6Ale4V mesh. The
physico-chemical conditions are established (Yerokhin et al., 1999),
exposed area during LSV was 4 cm2. The photocurrent measure-
promoting high growth rates (roughly 1 mm/min), oxide crystalli-
ments were repeated three times. The Incident Photon-to-Current
zation and incorporation of chemical species from the electrolyte
Efficiency (IPCE) was calculated using equation (2).
(Bayati et al., 2010; Franz et al., 2016; Mirelman et al., 2012). Due to
these features, since the ‘70s the PEO process has found industrial h*c I
applications to obtain oxide protective layers on aluminum and IPCEð%Þ ¼ * (2)
e P*l
magnesium (Yerokhin et al., 1999), and titanium alloys, mostly
Tie6Ale4V for biomedical implants (Liu et al., 2005; Nie et al., where h is the Planck constant [m2Kg s1], c is the speed of light [m
2000; Wei et al., 2008). s1], e is the electron charge [C], I is the steady-state photocurrent
In the present study, the effectiveness of the electrochemical density [A m2], P is the light intensity [W m2] and l is the inci-
photocatalysis employing immobilized TiO2 obtained by PEO was dent wavelength. The IPCE values were taken under polarization at
S. Murgolo et al. / Water Research 164 (2019) 114920 3

1.5 V vs SCE. The electrochemical surface area (ECSA) of the TiO2 detected in the employed wastewater and the investigation of TPs
obtained by PEO was estimated by cyclic voltammetry (Trasatti and (see next section) was carried out by an Ultimate 3000 UPLC System
Petrii, 1991), where consecutive potential cycles centered around (Thermo Fisher Scientific) equipped with an autosampler,
the open potential circuit were recorded in a potential temperature-controlled column compartment and UV detector as a
range 0.52 ÷ 0.42 V vs. SCE at five different scan rates. chromatographic system that was interfaced with a high-resolution
mass spectrometer, TripleTOF® 5600 þ System (AB Sciex) equipped
2.2. Organic pollutants and wastewater characteristics with a duo-spray ion source that was operated in electrospray (ESI)
mode in positive and negative ion modes. MS analysis was carried
Wastewater was collected at the outlet of the aerobic biological out by an information dependent analysis (IDA) method that in-
step (suspended biomass) from a WWTP of the area of Milan, Italy, cludes a survey scan in TOF-MS and, after background subtraction,
that includes grit and grease removal step, biological oxidation- the isolation and fragmentation in the collision cell of the four most
nitrification sections, denitrification, secondary sedimentation intense ions.
steps, disinfection. The physicochemical parameters were deter- The chromatographic separation of the organic mixtures was
mined according to Standard Methods (Rice et al., 2012) and are performed by injecting 500 mL samples, according to the so-called
listed in Table 1 together with the relevant figures of the selected large volume injection mode, and eluting them at 0.200 mL min1
WWTP. Wastewater samples were first filtered on paper filter to through a BEH C18 column, 2.1  150 mm, 1.7 mm, with a binary
remove coarse materials and then on 0.2 mm glass filter. Filtered gradient consisting of H2O/ACN 95/5, 0.1% HCOOH (A) and ACN,
wastewater was spiked with carbamazepine (Sigma-Aldrich) at 0.1% HCOOH (B). The large volume injection mode was used in
concentration of 100 mg L1 before running the photocatalytic ex- order to reach limit of detection of the order of few ng L1, namely
periments. The spiking concentration was chosen in order to obtain the concentration of CECs usually found in real wastewater efflu-
a good compromise between having a concentration high enough ents. The gradient started from 5% B that was held for 4.6 min and
to be able to detect a number of transformation products and low then was linearly increased to 80% in 18.5 min and to 100% in
enough to be close to real environmental conditions. further 5.5 min and the final composition was held for 5 min. At the
end of each run, the system was further rinsed for 6 min using the
final eluent composition and then returned to starting conditions
2.3. Experimental setup
and equilibrated for 5 min. The residual concentration of carba-
mazepine was obtained by MultiQuan 3.0.2 software (AB Sciex).
The process was carried out in a laboratory-scale 1 L tubular
photocatalytic reactor equipped with a 1 L buffer reservoir working
2.5. Suspects screening, non-target screening and transformation
in semi-batch mode under electrical polarization (bias) of the
products identification
expanded mesh (geometric surface area 327.5 cm2). The sketch of
the employed reactor is shown in Fig. S1 (Supplementary Material).
While targeted LCetandem MS (liquid chromatographyemass
The UV source consisted of a 30 W low-pressure Hg vapor UV-C
spectrometry) methods are specifically designed to analyze and
lamp emitting at 254 nm. Accordingly, four different configura-
quantify a set of previously defined target compounds, non-target
tions with the explained reactor were investigated as follow:
approaches make use of modern high speed and high resolution
TiO2þUV þ bias, TiO2þUV, TiO2þbias, and only UV. Control tests
mass spectrometers (HRMS) to gather as much information as
with the conventional suspended catalyst were carried out with a
possible about any detectable substance. All the collected IDA-MS
batch reactor (1 L) equipped with a 30 W low-pressure Hg vapor
data files, corresponding to the different reaction times for a spe-
lamp (Helios Italquartz). The employed conventional catalyst was
cific treatment, were processed for non-target screening using a
Degussa P25 (Evonik) TiO2, consisting of anatase and rutile crys-
data workflow in enviMass 3.4 software (Loos, 2018). Briefly, the
tallites with a ratio typically of 80:20, a surface area of 50 m2 g1
approach used for the interpretation and processing of data is as
and an average diameter of 30 nm. The radiance density flux at
follows: an initial enviMass peak picking step generates a list of
254 nm was measured through actinometry test using 10 mM uri-
ions with a corresponding retention time and peak intensity in each
dine (Sigma-Aldrich).
acquired file. The list of ions was subsequently reduced by replicate
sample intersection, removal of peaks which have also been
2.4. Chemical analysis detected in blank samples, isotope grouping and adduct grouping.
The obtained peak lists were then processed by SciexOS software
The monitoring at various reaction times of the residual con- for detecting significant trends as well as for identification of un-
centration of spiked carbamazepine, of the other compounds known compound, employing both the formula finder based on
isotopic pattern and library searching capabilities (LibraryView).
Table 1
Characteristics of the employed real secondary wastewater effluent.
Structure identification was then attempted based on high resolu-
tion MS-MS data (Detomaso et al., 2005). Moreover, linking these
parameter range value (min-max)
results to ChemSpider and Metlin, a more confident identification
Flow rate 432,000 m3 d1 of detected compounds in investigated samples was performed
Population equivalent 1,250,000 (Guijas et al., 2018). The peaks for which a typical trend of trans-
pH 7.3 ÷ 8
conductivity 720 ÷ 950 mS cm1
formation products was identified were further processed by R
COD 30 ÷ 60 mg L1 statistical environment, using a linkage analysis script, in order to
BOD5 <15 mg L1 obtain further information about the occurrence of possible TPs
TOC 12 ÷ 25 mg L1 (Schollee et al., 2018).
DOC 10 ÷ 22 mg L1
Cl 200 ÷ 260 mg L1
NeNO3 2 ÷ 4 mg L1 3. Results and discussion
NeNH4 <5 mg L1
Total N <10 mg L1 3.1. Characterization of nanostructured TiO2 meshes
Total P 0.9 ÷ 1.5 mg L1
SST <15 mg L1
As shown by the SEM micrograph of the surface of the TiO2
4 S. Murgolo et al. / Water Research 164 (2019) 114920

coating in Fig. 1a, the oxide layer is porous with an interconnected applied voltage and then it stabilizes at a maximum value of 65%.
sponge like morphology. The average surface porosity obtained by The previous characterizations were carried out on sacrificial
analyzing the SEM surface micrograph is about 10%. However, the samples obtained in the same experimental conditions as for the
cross-section SEM micrograph (inset of Fig. 1a) revealed that the expanded mesh employed in the reactor. From the practical point of
coatings are increasingly compact in depth. The GDOES depth view, linear sweep photovoltammetric measurements were carried
profiles of the titanium and oxygen elements shown in Fig. 1b out in situ by keeping the mesh in the reactor. This allows an easily
confirmed that the thickness of the TiO2 coating is roughly 2.5 mm. monitoring of the activity of the TiO2 catalyst grown on the
According to the XRD patterns shown in Fig. 1c, the as-prepared expanded mesh. Additionally, based on the photocurrent response
TiO2 films are crystalline in structure. The mass fraction of the of the mesh, the electrochemical polarization voltage to be applied
anatase and rutile phases calculated following Spurr (Spurr and at the catalyst during operation can be selected and possible phe-
Myers, 1957) are 58% and 42%, respectively. Based on the nomena of shielding of the TiO2 film due to the anodic deposition of
UVevisible spectra and the corresponding Kubelka-Munk conver- dispersed particulate or oxidation by-products can be detected. In
sion using the Tauc-plot method (Fig. S2) a band gap of 3.06 eV was Fig. S3, the resulting photocurrent response from 0 V to 5 V (cell
calculated, as expected considering the band gap of pure anatase voltage) is shown. In agreement with measurements carried out on
(3.2 eV) and pure rutile (3.02 eV). sacrificial samples, the photocurrent shows an initial sharp increase
The photoelectrochemical activity of TiO2 coating was assessed followed by a plateau.
by measuring the photocurrent of the films under electrical bias, i.e. The minimum surface area of the TiO2 catalyst employed in the
by linear voltammetry in dark and under UV-C irradiation from 0 V electrochemical photoreactor corresponded to 327.5 cm2, i.e. the
to 1.5 V (vs. SCE). The photocurrent density initially increases geometrical area of the expanded mesh itself. However, the real
steeply with the potential, and then it stabilizes on a well- surface area is expected be higher than the geometrical area of the
developed plateau at 0.174 mA cm2 and 0.96 V vs. SCE. The corre- titanium substrate due to the porous morphology. In order to es-
sponding IPCE, which is defined as the number of electrons timate the real surface area involved in the photoelectrochemical
generated by light in the external circuit divided by the number of process, the electrochemical surface area (ECSA) of the catalyst was
incident photons, more clearly represents the photo- measured (Trasatti and Petrii, 1991). The ECSA was calculated on
electrochemical activity of the coatings. In agreement with the the basis of the capacitive current due to the double-layer charging
photocurrent density, the IPCE initially linearly increases with the and discharging at the solideelectrolyte interface in the non-

Fig. 1. SEM micrographs, surface and cross-section (inset), of the TiO2 film (a), GDOES in-depth profile of Titanium and Oxygen elements into the TiO2 film (b), XRD pattern of the
TiO2 films (A ¼ Anatase; R ¼ Rutile; T ¼ Titanium substrate) (c) and IPCE of TiO2 films obtained by PEO as a function of the applied voltage (d).
S. Murgolo et al. / Water Research 164 (2019) 114920 5

faradaic region of the cyclic voltammograms. The resulting total suspended catalyst that is known to be difficult to be removed at
intensive capacitance of about 48.2 mF cm2 was compared to the the end of the water treatment process. In addition, the supported
experimental intrinsic specific capacitance, which was about catalyst was positively tested for 15 cycles showing that the per-
0.88 mF cm2, leading to a real surface area of 54.8 cm2 per cm2 of formance, based on pseudo-first order kinetic constants, remained
geometrical area. This corresponds to around 5.7 m2 g1, which is a in the range of 75e100% (Fig. 3a). After a prolonged use (several
reasonable value considering that TiO2 nanopowders have real tenths of reactions) the catalyst showed a reduced performance.
surface area of the order of 150 m2 g1. Therefore, the total ECSA of However, the initial performance was recovered by a chemical
the catalyst used in the electrochemical photoreactor was 1.795 m2. cleaning in dilute HCl solution or, eventually, by a re-anodization
Comparatively, the control tests in the batch reactor were carried process (Fig. 3b).
out using a total amount of conventional suspended catalyst of In order to get deeper insights about the effectiveness of the
100 mg L1, corresponding to a real surface area of 5 m2. novel catalyst, the comparison of the electrochemical photo-
catalysis with respect to both photocatalysis with the same sup-
ported TiO2 catalyst and photolysis was also performed for the CECs
3.2. Photocatalytic removal of CECs
naturally present in the secondary wastewater effluent. An
analytical screening was carried out, as described in section 2.5,
The degradation of carbamazepine, which was initially spiked at
aimed at identify as many as possible CECs. The procedure led to the
100 mg L1 in the wastewater effluent, by the conventional photo-
detection of 201 compounds present in the secondary wastewater
catalysis with suspended Degussa P25 and by electrochemical TiO2
effluent employed for the photocatalysis tests and 51 of them were
photocatalysis is depicted in Fig. 2. Results show that with the two
above mentioned photocatalytic processes the carbamezapine was
completely removed within 45 min. The degradation by the con-
ventional suspended catalyst showed a slightly faster removal in
the first stage of the reaction, namely between 5 and 30 min. The
beneficial effect of applying a polarization potential (bias) to the
supported TiO2 catalyst is also evident from Fig. 2 since the reaction
performed without any bias showed a much slower carbamazepine
degradation being about 20% the residual carbamazepine after
90 min of reaction. Also, reaction performed with conventional and
electrochemical photocatalysis shows a faster carbamazepine
removal than photolytic process. A more equitable comparison
could be obtained where the results of actinometry test in both
conventional and electrochemical photocatalysis conditions are
considered. In particular, it was revealed that in case of electro-
chemical photocatalysis, the presence of the mesh introduces a
shielding effect of almost 66% remaining an amount of radiance
density flux as 0.08 W cm2 while in the case of conventional
photocatalysis, the shielding effect was measured as 6.74% corre-
sponding to a flux of 0.15 W cm2. Overall, the results displayed in
Fig. 2 demonstrate the effectiveness of electrochemical photo-
catalysis with supported TiO2 catalyst. Specifically, even though the
performance of the electrochemical TiO2 photocatalysis showed to
be similar to the photocatalysis with the conventional suspended
TiO2, the former process has the advantage of avoiding using a

Fig. 2. Degradation of carbamazepine by photoelectrocatalysis and conventional Fig. 3. Replicates of carbamazepine degradation in real secondary effluent wastewater
photocatalysis (Degussa P25) performed in real secondary effluent wastewater. Error by photoelectrocatalysis (a) and aging effect of the nanostructured TiO2 meshes along
bars represent the standard deviation obtained for three replicates. prolonged use and recovery of the initial performance by re-anodization (b).
6 S. Murgolo et al. / Water Research 164 (2019) 114920

also successfully associated to CECs on the basis of the positive photocatalysis is effective in removing not only the spiked carba-
match of both the isotopic cluster of molecular ion and the product mazepine at 100 mg L1 but, more importantly, a high number of
ion spectrum with that of the available MS library. The decay of the CECs present in the real secondary wastewater effluent.
51 identified CECs was found to follow first-order kinetics and the
distribution of their kinetic constants are shown in Fig. 4a. Boxplots
3.3. Minimization of transformation products
built with all obtained kinetic constant values show the higher
performance of the electrochemical photocatalysis with respect to
Beside the degradation of CECs present in the employed sec-
the experiment where the bias was off (TiO2þUV) as well as to the
ondary wastewater effluent, it is also important to assess the TPs
photolysis. Specifically, the median values of the 51 kinetic con-
arising from the treatment by electrochemical photocatalysis with
stants resulted to be 0.044, 0.023 and 0.034 min1 for the
the novel supported TiO2 catalyst with respect to both photolysis
TiO2þBias þ UV, TiO2þUV and UV, respectively. A similar trend
and photocatalysis with the same supported catalyst. The
with the highest values for the TiO2þBias þ UV process was found
employed analytical protocol for the screening of non-target or-
when considering the mean kinetic constant value (0.048, 0.035
ganics allowed detecting compounds not present in the secondary
and 0.043 min1, respectively) and the 25e75% percentile range
effluent at the beginning of the UV-based experiments that could
used for building the boxplots (0.02e0.069, 0.007e0.053 and
be thus confidently associated to TPs. Using the same protocol, the
0.009e0.059, respectively).
detection of compounds initially present in the employed waste-
From the first-order kinetic constants, it can be easily obtained
water and increasing along reaction time was also accomplished.
the half-life times of each suspect CECs. All the results were plotted
Consequently, the latter compounds were associated to TPs, too. By
as boxplots and the results showed, once again, the superior per-
a careful inspection of the detected compounds as output of the
formance of the electrochemical photocatalysis (Fig. 4b). Specif-
aforementioned analytical protocol, 194 TPs were rationally iden-
ically, a mean half-life time of 28.1 min was calculated for the
tified (Table S2). The detected TPs showed to follow three different
TiO2þBias þ UV process with respect to 80 and 62.3 min for the
types of time profiles, namely (i) a bell-shape trend, (ii) slight in-
TiO2þUV and UV treatments, respectively. In addition, from the
crease followed by a steady time-profile, (iii) a constant increase
first-order kinetic constants the Electrical Energy per Order of
along reaction time. TPs with the latter two time-profiles were thus
magnitude of removal (EEO), expressed in kWh m3, can be calcu-
accumulating in the reaction mixture.
lated using equation (1) (Murgolo et al., 2017):
Indeed, a number of TPs were surely arising from the organic
! matter contained in the real wastewater rather than from CECs.
. 38:4 x UV power ðkWÞ
EEO kWh m3 ¼   (3) Therefore, considering the high number of TPs detected, the com-
V ðLÞ x k min1 parison of the three UV-based processes was performed based on
average peak intensity. In Fig. 5a are depicted the results of the
where k is the first-order rate constant (min1) for the disappear- 194 TPs for the three UV-based treatments at both the final reaction
ance or the target CECs, V is the water volume to be treated and UV time (90 min) and the middle reaction time (45 min) in comparison
power is obtained from the employed UV lamp. As expected, the with the initial wastewater. The boxplots of Fig. 5a clearly show the
obtained results (Fig. 4c) show the lower energy requirements of higher performance of the electrochemical photocatalysis with
the electrochemical photocatalysis being the EEO values 46.8, 133 respect to both the experiments in which the bias was off (TiO2-
and 103.5 kWh m3, for the for the TiO2þBias þ UV process with þUV) and the photolysis. Specifically, the mean peak area for
respect to the TiO2þUV and UV treatments, respectively. Indeed, TiO2þBias þ UV at 45 min reaction time resulted to be almost the
the obtained EEO values when translated into electrical operating half those of the other two treatments (183500, 31200 and 35900
costs would lead to values at least one order of magnitude higher for the TiO2þBias þ UV, TiO2þUV and UV, respectively). Results
than those calculated for CECs removal at full scale WWTPs. It has obtained at 90 min reaction time show that such a difference be-
to be taken into account that the employed experimental system is tween TiO2þBias þ UV and the other two processes further
a small lab-scale reactor and consequently its electrical operating increased being the mean peak area one-third of the other two
cost cannot be compared with that of full-scale systems. reactions. It is worth noting that the mean peak area value obtained
Overall, the above-described results demonstrate that the novel by TiO2þBias þ UV after 90 min is only 67% of that of the initial
supported TiO2 catalyst employed for the electrochemical wastewater (110000 with respect to 163000) because only the

Fig. 4. Box-whisker plots of kinetic constants for the 51 suspect CECs identified by the analytical screening procedure (Table S1) during degradation by photolysis, photocatalysis
with supported TiO2 and electrochemical photocatalysis (TiO2þBias þ UV) in real secondary effluent wastewater. Boxplots represent the distance between the first and third
quartiles while whiskers are set as the most extreme (lower and upper) data point not exceeding 1.5 times the quartile range from the median. Kinetic constants outside such a
range are outliers.
S. Murgolo et al. / Water Research 164 (2019) 114920 7

Fig. 5. Box-whisker plots of peak area for the 194 TPs (a) and for all 395 compounds (194 TPs þ 201 compounds present in the secondary wastewater effluent) detected by the
analytical screening procedure (Table S2) during degradation by photolysis, photocatalysis with supported TiO2 and electrochemical photocatalysis (TiO2þBias þ UV) in real
secondary effluent wastewater at 45 and 90 min reaction time. Boxplots represent the distance between the first and third quartiles while whiskers are set as the most extreme
(lower and upper) data point not exceeding 1.5 times the quartile range from the median. Values outside such a range are outliers.

detected 194 TPs were considered where, most of them, were not photocatalysis (Fig. 6). Specifically, four transformation reactions
present in the initial wastewater. When both the detected 194 TPs (addition of 2 atoms of oxygen, addition of 3 atoms of oxygen,
and the 201 compounds present in the secondary wastewater hydroxylation/N or S-oxidation/epoxidation, alcohol to carboxylic
effluent employed for the photocatalysis tests (51 of which were acid) were found to have, respectively, 8, 5, 7 and 10 couples of
associated to CECs) were considered, different results were ob- parent/TP that could be positively associated to them. The oxidative
tained (Fig. 5b). Specifically, the mean peak area value of the initial displacement of amine reaction was found to be the transformation
wastewater was similar to those of the TiO2þUV and UV treatments reaction with the highest number (11) of matches. This reaction
at both 45 and 90 min but the correspondent values of the occurred on nitrogen-containing compounds present in the
TiO2þBias þ UV treatment were consistently lower. In addition, wastewater matrix and the UV light made such a reaction likely to
Fig. 5b shows that, even though several TPs were detected, after just occur. It is worth noting that beside the hydration reaction also the
45 min of reaction the three investigated processes were able to de-hydration reaction was found to have 6 positive matches. It
lead an average peak area of all detected compounds lower than the follows that both the hydration and de-hydration occur during UV-
initial value, being that of the TiO2þBias þ UV treatment the lowest. based processes with real wastewater due to the chemical
This demonstrates that the TiO2þBias þ UV treatment is much composition of the real water matrix and to the combination of UV
more efficient than both TiO2þUV and UV in minimizing the in- light and temperature increase as a consequence of UV lamp energy
tensity of the organics in the real wastewater. Such a better per- dissipation. In addition, a relevant number of positive matches
formance was more pronounced at higher reaction time reaching were also observed for reaction likely due to photolysis, namely di-
60% reduction of mean peak area of TPs at 90 min of reaction. demethyl or de-ethylation (9 matches), demethylation and de-
acetylation (6 matches for each reaction), dealkylation (5 matches).
The linkage analysis was also carried out using, as parent com-
3.4. Identification of transformation products and carbamazepine pounds, the list of 51 suspect CECs identified within the 201 com-
degradation pathway pounds detected in the initial wastewater. Results (Fig. 6, red lines)
showed that the distribution of detected transformations was
In order to get insights about the occurrence of TPs arising from similar to the previous linkage analysis. Specifically, the highest
possible transformation reaction the linkage analysis was number of matches were obtained for the alcohol to carboxylic acid
employed (Scholle e et al., 2018). The linkage analysis is based on reaction, de-hydration, dealkylation (4 matches for each trans-
the fact that a transformation reaction is associated to an accurate formation) and hydroxylation/N or S-oxidation/epoxidation, hy-
mass difference between the parent compound and the TP. As the dration, dealkylation (3 matches).
analytical screening was performed by high resolution-mass In addition, as carbamazepine was spiked in the real wastewater
spectrometry for all detected compounds the accurate mass was at beginning of the reaction, the couples of parent/TP detected by
obtained (Tables S1 and S2). A list of possible transformation re- the linkage analysis were inspected in order to find those that
actions was built on the basis of reactions likely to occur during UV- matched the known TPs of carbamazepine. Six carbamazepine TPs
based processes and each reaction is consequently associated to an were positively detected that allowed to drawn the degradation
accurate mass difference (Table 2). The list of parent compounds pathway depicted in Fig. 7. It is worth noting that all the TPs of Fig. 6
detected in the initial wastewater (Table S1) was then compared were positively identified on the basis of accurate mass of molec-
with the list of detected TPs (Table S2) in order to obtain the couples ular ion, elemental composition using the isotopic cluster of mo-
of parent/TP whose accurate mass difference match one of those lecular ion and MS-MS spectra. Therefore, the identification level of
listed in Table 2 within the set mass tolerance. As expected, results such TPs was 2 (probable structure) according to Schymanski et al.,
showed that the highest number of transformations were those (2014). The TP250 (Fig. 7) was rationalized to derive from TP268
compatible with oxidative reactions likely to occur during
8 S. Murgolo et al. / Water Research 164 (2019) 114920

Table 2
List of transformation reactions considered for linkage analysis with the indication of the exact mass difference and formula difference between the parent
compound and transformation product as a result of the transformation reaction.

Transformation reaction Formula difference Exact m/z difference

Glucuronidation þC6H8O6 176.032


Sulfation þSO3 79.9568
Addition of 3 atoms of oxygen þ3O 47.9847
Acetylation þC2H2O 42.0106
Addition of 2 atoms of oxygen þ2O 31.9898
Methyl to carboxylic acid, amine to nitro -2Hþ2O 29.9741
Hydration þH2O 18.0106
Hydroxylation, N or S-oxidation, epoxidation þO 15.9949
Oxidative deamination -NH3þ2O 14.9632
Alcohol to carboxylic acid or primary amine to nitro -2H þ O 13.9792
Oxidative displacement of amine -NH þ OH 1.9918
Deamination to ketone -NH3þO 1.0317
Demethylation -CH2 14.0157
Dehydration -H2O 18.0106
Di-demethyl or de-ethylation -C2H4 28.0313
Dealkylation -C2H6 30.047
De-acetylation -C2H2O 42.0106
Decarboxylation -CO2 43.9898
Loss of nitro group -NO2þH 44.9851
De-sulfation -SO3 79.9568
De-glucuronidation -C6H8O6 176.032

Fig. 6. Radar plot of the different transformation reactions detected by linkage analysis considering as parent compounds all the 201 compounds listed in Table S1 (blue line) or the
51 suspect CECs detected within such a list (red line). The scale, indicated on the gray concentric circles, shows the number of times each transformation reaction was detected. (For
interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

(and not from carbamazepine by the alcohol to carboxylic acid or forming low-molecular weight organic acids as final degradation
primary amine to nitro reaction that was also detected by the products that are not detectable by the employed chromatographic
linkage analysis) by a hydration reaction as this TP was also column being necessary to use ion chromatography (Mascolo et al.,
detected in the initial wastewater. TP223 was found by the linkage 2005). It worth noting that the formation of such further TPs would
analysis as parent of the TP179 (acridine) and it was rationalized to not be easily detectable in the employed real effluent since car-
derive from carbamazepine as a result of multiple transformations. boxylic acids could exists as coming from other processes and other
In Fig. S4 the time-profiles of the TPs of Fig. 6 are depicted, organic compounds.
showing that the electrochemical TiO2 photocatalysis allowed a By inspecting the couples of parent/TP detected by the linkage
higher minimization of their abundances with respect to both analysis it was also possible to find some compounds that under-
photolysis and photocatalysis with supported TiO2. These results went transformation reactions. These compounds are not included
confirmed what above reported about the better performance of in the list of suspect CECs and therefore are likely to be compounds
the investigated process. It is worth noting that the time-profiles present in the water matrix of the employed real wastewater. The
show that at higher reaction time the TPs disappear likely reactions detected by the linkage analysis were, most of the times,
S. Murgolo et al. / Water Research 164 (2019) 114920 9

Fig. 7. Proposed degradation pathway of carbamazepine obtained on the basis of the transformation reactions detected by the linkage analysis. TPs are named with their nominal
mass.

oxidation reactions that are consistent with the investigated pro- demonstrated that the supported catalyst can be easily reused
cesses. A typical example is related to the parent compound of without losing efficiency. The fate of 194 metabolites formed during
accurate mass 141.1154 ([MþH]þ ¼ 142.1218) whose elemental the investigated treatments was assessed, observing a higher per-
composition was determined to be C8H15NO. This compound is formance of the electrochemical photocatalysis with respect to
likely to be 1-piperidinoacetone even though two other chemical both the experiments in which the bias was off (TiO2þUV) and the
structures are possible, namely tropine and azonan-2-one. For the photolysis. Results show that non-target evaluation can give addi-
parent compound the linkage analysis detected two oxidative tional information to assist process evaluation beyond the analysis
transformation products as depicted in the degradation pathway of of single “target” micropollutants. The suspect search allows a
Fig. 8 where the chemical structure of two other possible parent wider spectrum of micropollutants to be monitored (e.g., due to a
compounds are shown, too. The chemical structure could not be decrease in treatment performance) and can be used to annotate
identified unequivocally due to lack of MS-MS spectra and non-target peak lists to identify the peaks which are already known
authentic standard. Overall, the linkage analysis demonstrated to prior to further non-target evaluation. Finally, the possible scale-up
be a useful tool for selecting the TPs to be further investigated of the proposed water treatment could take advantage from the fact
among the full list of TPs. that the catalyst is prepared by a technique already in use in the
industry of surface treatments. As far as information about oper-
4. Conclusions ating costs is concerned, it can be stated that the investigated
process is less energy demanding of the other two tested processes.
The application of an electrochemical photocatalytic process However, as the employed experimental reactor is a small lab-scale
based on the use of an innovative photoactive coating grown system, the obtained EEO values when translated into electrical
directly on titanium meshes was evaluated for the degradation of operating costs would lead to values at least one order of magni-
pharmaceuticals in real secondary wastewater effluent. Carba- tude higher than those calculated for CECs removal at full scale
mazepine and other CECs were more rapidly phototransformed WWTPs. It follows that the new investigated process still needs
compared to only photolysis, regardless the water matrix. It was many improvements to be a reality in scale-up.

Fig. 8. Proposed degradation pathway of 1-piperidinoacetone obtained on the basis of the transformation reactions detected by the linkage analysis. The two other possible
chemical structures, namely tropine and azonan-2-one, are also shown.
10 S. Murgolo et al. / Water Research 164 (2019) 114920

Acknowledgments national policy on reducing micropollutants in treated wastewater. Environ. Sci.


Technol. 48 (21), 12500e12508.
Loos, M., 2018. EnviMass version 3.5 LC-HRMS trend detection workflow - R
The authors acknowledge Dr. Gian Luca Chiarello, Department of package.
Chemistry, University of Milano, for the measurement of the Loos, R., Carvalho, R., Anto nio, D.C., Comero, S., Locoro, G., Tavazzi, S., Paracchini, B.,
emission spectrum of the lamp. Ghiani, M., Lettieri, T., Blaha, L., Jarosova, B., Voorspoels, S., Servaes, K.,
Haglund, P., Fick, J., Lindberg, R.H., Schwesig, D., Gawlik, B.M., 2013. EU-wide
monitoring survey on emerging polar organic contaminants in wastewater
Appendix A. Supplementary data treatment plant effluents. Water Res. 47 (17), 6475e6487.
Luo, Y., Guo, W., Ngo, H.H., Nghiem, L.D., Hai, F.I., Zhang, J., Liang, S., Wang, X.C.,
2014. A review on the occurrence of micropollutants in the aquatic environ-
Supplementary data to this article can be found online at ment and their fate and removal during wastewater treatment. Sci. Total En-
https://doi.org/10.1016/j.watres.2019.114920. viron. 473e474, 619e641.
Mascolo, G., Lopez, A., Detomaso, A., Lovecchio, G., 2005. Ion chromatography-
electrospray mass spectrometry for the identification of low-molecular-
References weight organic acids during the 2,4-dichlorophenol degradation.
J. Chromatogr. A 1067 (1e2), 191e196.
Andronic, L., Isac, L., Miralles-Cuevas, S., Visa, M., Oller, I., Duta, A., Malato, S., 2016. Miklos, D.B., Remy, C., Jekel, M., Linden, K.G., Drewes, J.E., Hübner, U., 2018. Evalu-
Pilot-plant evaluation of TiO2 and TiO2-based hybrid photocatalysts for solar ation of advanced oxidation processes for water and wastewater treatment e a
treatment of polluted water. J. Hazard Mater. 320, 469e478. critical review. Water Res. 139, 118e131.
Balest, L., Mascolo, G., Di Iaconi, C., Lopez, A., 2008. Removal of endocrine disrupter Mirelman, L.K., Curran, J.A., Clyne, T.W., 2012. The production of anatase-rich
compounds from municipal wastewater by an innovative biological technology. photoactive coatings by plasma electrolytic oxidation. Surf. Coat. Technol. 207,
Water Sci. Technol. 58 (4), 953e956. 66e71.
Bayati, M.R., Moshfegh, A.Z., Golestani-Fard, F., 2010. In situ growth of Murgolo, S., Yargeau, V., Gerbasi, R., Visentin, F., El Habra, N., Ricco, G., Lacchetti, I.,
vanadiaetitania nano/micro-porous layers with enhanced photocatalytic per- Carere, M., Curri, M.L., Mascolo, G., 2017. A new supported TiO2 film deposited
formance by micro-arc oxidation. Electrochim. Acta 55 (9), 3093e3102. on stainless steel for the photocatalytic degradation of contaminants of
Bestetti, M., Franz, S., Cuzzolin, M., Arosio, P., Cavallotti, P.L., 2007. Structure of emerging concern. Chem. Eng. J. 318, 103e111.
nanotubular titanium oxide templates prepared by electrochemical anodization Nakata, K., Fujishima, A., 2012. TiO2 photocatalysis: design and applications.
in H2SO4/HF solutions. Thin Solid Films 515 (13), 5253e5258. J. Photochem. Photobiol. C Photochem. Rev. 13 (3), 169e189.
Borges, M., García, D., Herna ndez, T., Ruiz-Morales, J., Esparza, P., 2015. Supported Nie, X., Leyland, A., Matthews, A., 2000. Deposition of layered bioceramic hy-
photocatalyst for removal of emerging contaminants from wastewater in a droxyapatite/TiO2 coatings on titanium alloys using a hybrid technique of
continuous packed-bed photoreactor configuration. Catalysts 5 (1), 77. micro-arc oxidation and electrophoresis. Surf. Coat. Technol. 125 (1), 407e414.
Byrne, C., Subramanian, G., Pillai, S., 2017. Recent Advances in Photocatalysis for Ochiai, T., Fujishima, A., 2012. Photoelectrochemical properties of TiO2 photo-
Environmental Applications. catalyst and its applications for environmental purification. J. Photochem.
Castiglioni, S., Davoli, E., Riva, F., Palmiotto, M., Camporini, P., Manenti, A., Photobiol. C Photochem. Rev. 13 (4), 247e262.
Zuccato, E., 2018. Mass balance of emerging contaminants in the water cycle of Pagano, M., Lopez, A., Volpe, A., Mascolo, G., Ciannarella, R., 2008. Oxidation of
a highly urbanized and industrialized area of Italy. Water Res. 131, 287e298. nonionic surfactants by Fenton and H2O2/UV processes. Environ. Technol. 29
Cates, E.L., 2017. Photocatalytic water treatment: so where are we going with this? (4), 423e433.
Environ. Sci. Technol. 51 (2), 757e758. Petronella, F., Diomede, S., Fanizza, E., Mascolo, G., Sibillano, T., Agostiano, A.,
Comparelli, R., Cozzoli, P.D., Curri, M.L., Agostiano, A., Mascolo, G., Lovecchio, G., Curri, M.L., Comparelli, R., 2013. Photodegradation of nalidixic acid assisted by
2004. Photocatalytic degradation of methyl-red by immobilised nanoparticles TiO2 nanorods/Ag nanoparticles based catalyst. Chemosphere 91 (7), 941e947.
of TiO2 and ZnO. Water Sci. Technol. 49 (4), 183e188. Petronella, F., Fanizza, E., Mascolo, G., Locaputo, V., Bertinetti, L., Martra, G.,
Del Moro, G., Mancini, A., Mascolo, G., Di Iaconi, C., 2013. Comparison of UV/H2O2 Coluccia, S., Agostiano, A., Curri, M.L., Comparelli, R., 2011. Photocatalytic ac-
based AOP as an end treatment or integrated with biological degradation for tivity of nanocomposite catalyst films based on nanocrystalline metal/semi-
treating landfill leachates. Chem. Eng. J. 218, 133e137. conductors. J. Phys. Chem. C 115 (24), 12033e12040.
Del Moro, G., Prieto-Rodríguez, L., De Sanctis, M., Di Iaconi, C., Malato, S., Rice, E., Baird, R., Eaton, A., Clesceri, L., 2012. Standard Methods for the Examination
Mascolo, G., 2016. Landfill leachate treatment: comparison of standalone of Water and Wastewater.
electrochemical degradation and combined with a novel biofilter. Chem. Eng. J. Rizzo, L., Malato, S., Antakyali, D., Beretsou, V.G., Ðoli c, M.B., Gernjak, W., Heath, E.,
288, 87e98. Ivancev-Tumbas, I., Karaolia, P., Lado Ribeiro, A.R., Mascolo, G., McArdell, C.S.,
Detomaso, A., Mascolo, G., Lopez, A., 2005. Characterization of carbofuran photo- Schaar, H., Silva, A.M.T., Fatta-Kassinos, D., 2019. Consolidated vs new advanced
degradation by-products by liquid chromatography/hybrid quadrupole time-of- treatment methods for the removal of contaminants of emerging concern from
flight mass spectrometry. Rapid Commun. Mass Spectrom. 19 (15), 2193e2202. urban wastewater. Sci. Total Environ. 655, 986e1008.
Dong, H., Zeng, G., Tang, L., Fan, C., Zhang, C., He, X., He, Y., 2015. An overview on Schollee, J.E., Bourgin, M., von Gunten, U., McArdell, C.S., Hollender, J., 2018. Non-
limitations of TiO2-based particles for photocatalytic degradation of organic target screening to trace ozonation transformation products in a wastewater
pollutants and the corresponding countermeasures. Water Res. 79, 128e146. treatment train including different post-treatments. Water Res. 142, 267e278.
Fagan, R., McCormack, D.E., Dionysiou, D.D., Pillai, S.C., 2016. A review of solar and Schro 
€der, P., Helmreich, B., Skrbi 
c, B., Carballa, M., Papa, M., Pastore, C., Emre, Z.,
visible light active TiO2 photocatalysis for treating bacteria, cyanotoxins and Oehmen, A., Langenhoff, A., Molinos, M., Dvarioniene, J., Huber, C.,
contaminants of emerging concern. Mater. Sci. Semicond. Process. 42, 2e14. Tsagarakis, K.P., Martinez-Lopez, E., Pagano, S.M., Vogelsang, C., Mascolo, G.,
Franz, S., Perego, D., Marchese, O., Bestetti, M., 2015. Photoelectrochemical 2016. Status of hormones and painkillers in wastewater effluents across several
advanced oxidation processes on nanostructured TiO2 catalysts: decolorization European statesdconsiderations for the EU watch list concerning estradiols
of a textile azo-dye. J. Water Chem. Technol. 37 (3), 108e115. and diclofenac. Environ. Sci. Pollut. Control Ser. 23 (13), 12835e12866.
Franz, S., Perego, D., Marchese, O., Lucotti, A., Bestetti, M., 2016. Photoactive TiO2 Schymanski, E.L., Jeon, J., Gulde, R., Fenner, K., Ruff, M., Singer, H.P., Hollender, J.,
coatings obtained by Plasma Electrolytic Oxidation in refrigerated electrolytes. 2014. Identifying small molecules via high resolution mass spectrometry:
Appl. Surf. Sci. 385, 498e505. communicating confidence. Environ. Sci. Technol. 48 (4), 2097e2098.
Giannakis, S., Gamarra Vives, F.A., Grandjean, D., Magnet, A., De Alencastro, L.F., Sousa, J.C.G., Ribeiro, A.R., Barbosa, M.O., Pereira, M.F.R., Silva, A.M.T., 2018. A review
Pulgarin, C., 2015. Effect of advanced oxidation processes on the micro- on environmental monitoring of water organic pollutants identified by EU
pollutants and the effluent organic matter contained in municipal wastewater guidelines. J. Hazard Mater. 344, 146e162.
previously treated by three different secondary methods. Water Res. 84, Spurr, R.A., Myers, H., 1957. Quantitative analysis of anatase-rutile mixtures with an
295e306. X-ray diffractometer. Anal. Chem. 29 (5), 760e762.
Guijas, C., Montenegro-Burke, J.R., Domingo-Almenara, X., Palermo, A., Warth, B., Stalter, D., Magdeburg, A., Weil, M., Knacker, T., Oehlmann, J., 2010. Toxication or
Hermann, G., Koellensperger, G., Huan, T., Uritboonthai, W., Aisporna, A.E., detoxication? In vivo toxicity assessment of ozonation as advanced wastewater
Wolan, D.W., Spilker, M.E., Benton, H.P., Siuzdak, G., 2018. METLIN: a technology treatment with the rainbow trout. Water Res. 44 (2), 439e448.
platform for identifying knowns and unknowns. Anal. Chem. 90 (5), Ternes, T.A., 1998. Occurrence of drugs in German sewage treatment plants and
3156e3164. rivers1Dedicated to Professor Dr. Klaus Haberer on the occasion of his 70th
€ bel, A., McArdell, C.S., Ternes, T., Siegrist, H., 2005.
Joss, A., Keller, E., Alder, A.C., Go birthday.1. Water Res. 32 (11), 3245e3260.
Removal of pharmaceuticals and fragrances in biological wastewater treatment. Tran, N.H., Reinhard, M., Gin, K.Y.-H., 2018. Occurrence and fate of emerging con-
Water Res. 39 (14), 3139e3152. taminants in municipal wastewater treatment plants from different
Kasprzyk-Hordern, B., Dinsdale, R.M., Guwy, A.J., 2009. The removal of pharma- geographical regions-a review. Water Res. 133, 182e207.
ceuticals, personal care products, endocrine disruptors and illicit drugs during Trasatti, S., Petrii, O.A., 1991. Real Surface Area Measurements in Electrochemistry,
wastewater treatment and its impact on the quality of receiving waters. Water p. 711.
Res. 43 (2), 363e380. Wei, D., Zhou, Y., Jia, D., Wang, Y., 2008. Chemical treatment of TiO2-based coatings
Liu, F., Song, Y., Wang, F., Shimizu, T., Igarashi, K., Zhao, L., 2005. Formation char- formed by plasma electrolytic oxidation in electrolyte containing nano-HA,
acterization of hydroxyapatite on titanium by microarc oxidation and hydro- calcium salts and phosphates for biomedical applications. Appl. Surf. Sci. 254
thermal treatment. J. Biosci. Bioeng. 100 (1), 100e104. (6), 1775e1782.
Logar, I., Brouwer, R., Maurer, M., Ort, C., 2014. Cost-benefit analysis of the Swiss Yerokhin, A.L., Nie, X., Leyland, A., Matthews, A., Dowey, S.J., 1999. Plasma
S. Murgolo et al. / Water Research 164 (2019) 114920 11

electrolysis for surface engineering. Surf. Coat. Technol. 122 (2), 73e93. Zwiener, C., Seeger, S., Glauner, T., Frimmel, F., 2002. Metabolites from the
Zhou, J.L., Zhang, Z.L., Banks, E., Grover, D., Jiang, J.Q., 2009. Pharmaceutical residues biodegradation of pharmaceutical residues of ibuprofen in biofilm reactors and
in wastewater treatment works effluents and their impact on receiving river batch experiments. Anal. Bioanal. Chem. 372 (4), 569e575.
water. J. Hazard Mater. 166 (2), 655e661.

You might also like