You are on page 1of 12

Water Air Soil Pollut (2017) 228:339

DOI 10.1007/s11270-017-3532-0

Performance Analysis of Photolytic, Photocatalytic,


and Adsorption Systems in the Degradation of Metronidazole
on the Perspective of Removal Rate and Energy Consumption
Neghi N & Mathava Kumar

Received: 6 January 2017 / Accepted: 6 August 2017


# Springer International Publishing AG 2017

Abstract The efficiency of the following systems: pho- of MNZ via aliphatic carboxylic acid compounds in the
tolysis (UV-C only), photocatalysis with titanium-dioxide photocatalytic system. Overall, the photocatalytic system
(UV-C/TiO2), photocatalysis with granular-activated car- seems to be an energy-efficient treatment option for the
bon (UV-C/GAC), and by adsorption on GAC, was removal of MNZ and similar other micropollutants.
assessed under different initial contaminant concentrations,
i.e., 0.1–100 mg L−1. The experiments were conducted in a Keywords Antibiotics . Oxidative degradation
batch photocatalytic reactor (1.9 L and 32 W UV power). pathway . Metronidazole . Photocatalysis . Wastewater
It was found that UV-C/TiO2 and UV-C/GAC systems treatment . Energy consumption analysis
showed fairly equal removal efficiencies under lower
MNZ concentrations (0.1–5 mg L−1) compared to higher
concentrations at similar catalyst loading of 2.5 g L−1. A
1 Introduction
decline in removal rate (based on first-order reaction) was
observed with respect to increase in initial MNZ concen-
In the recent years, traces of human and animal pharma-
tration in all systems. MNZ removal by adsorption on
ceuticals have been observed in several aqueous systems.
GAC was much lesser compared to UV-C only, UV-C/
Moreover, the effects of pharmaceuticals to various envi-
TiO2, and UV-C/GAC systems. The adsorption data well
ronmental species have been discussed throughout the
correlated with the Freundlich model indicated that the
world. However, pharmaceuticals are one of the inevita-
adsorption was on the heterogeneous surface of the cata-
ble concomitants of everyday activity. Several metric tons
lyst. The effectiveness of the systems were evaluated by
of pharmaceuticals are used annually for controlling dis-
calculating electrical energy consumed per order (EEO).
eases and as feed supplements in poultries, aquaculture
The lowest EEO value was found to be for UV-C/TiO2
farms, animal husbandries, etc. Antibiotics, one of the
(0.03 kWh m−3 order−1) for the degradation of 0.1 mg L−1
major classes of pharmaceuticals, are designed to stop/kill
of MNZ compared to UV-C/GAC (0.06 kWh m−3 or-
the activity of disease-producing microorganisms, i.e.,
der−1), UV-C only (0.15 kWh m−3 order−1), and adsorption
microbial biochemical processes. The release of antibi-
(0.44 kWh m−3 order−1). The total organic carbon and
otics and/or their intermediates in the environment could
nitrogen ion analyses have confirmed the mineralization
produce detrimental effects to non-targeted indigenous
organisms such as cyanobacteria, algae, crustaceans, fish,
N. N : M. Kumar (*) etc. (Guo et al. 2016). The major sources of antibiotic
Environmental and Water Resources Engineering Division, release into various environmental compartments include
Department of Civil Engineering, IIT Madras, Chennai 600036,
India
hospitals, antibiotic formulation/production units, munic-
e-mail: mathav@iitm.ac.in ipal wastewater, households, meat processing units, ma-
e-mail: mathavakumar@gmail.com nure from husbandries, and sewage sludge (Coutu et al.
339 Page 2 of 12 Water Air Soil Pollut (2017) 228:339

2013; Lien et al. 2016; Perez et al. 2015). In water and biologically inert, photoactive, inexpensive, and
systems, antibiotics were detected in the range of nano- non-toxic and (2) the hydroxyl radical (HO·) produced
grams per liter to micrograms per liter (Asha and Kumar in the system has strong oxidation potential (+ 2.80 V),
2015); however, higher concentration levels in the waste- which is non-selective in rapidly oxidizing (rate constants
waters released from pharmaceutical production and mu- in the order of 106–1010 M−1 s−1) most organic com-
nicipal wastewater treatment units were reported pounds (Petrovic et al. 2011; Andreozzi et al. 1999).
(Kimosop et al. 2016). The exposure of indigenous or- The mechanism of UV-C/TiO2 in the degradation of
ganisms to antibiotics (especially bacteria) could develop organic micropollutants (R) including MNZ is shown in
antibiotic-resistant/antimicrobial-resistant (AMR) genes, Eqs. (1)–(5).
which has the capability to produce detrimental effect to
TiO2 þ hv ð254 nmÞ→e−CB þ hþ
VB ð1Þ
the humankind (Lupo et al. 2012). Therefore, it is of
utmost concern to stop the entry of antibiotics into the
environment by introducing policy decisions and to de- H 2 O þ hþ
VB →HO• þ H
þ
ð2Þ
sign suitable treatment systems for antibiotics removal.
Metronidazole (C6H9N3O3, MNZ) is one of the pri-
ority antibiotics, which has been used for controlling O2 þ e−CB →O−2 • ð3Þ
infections caused by anaerobic bacteria, protozoa, and
bacteroides (Bendesky et al. 2002). MNZ has a wide-
spread application in human pharmaceuticals and at the R þ hþ
VB →R

ð4Þ
same time, it has been used as an additive in poultry and
fish farms. The cytotoxic intermediates released from
MNZ after ingestion have the potential to damage DNA, R þ HO•→R• →CO2 þ H2 O ð5Þ
protein, and membrane of the disease-producing organ-
isms. It was reported that MNZ is genotoxic to humans Where e−CB ; hþ
VB are electrons and holes produced in
and it can also produce mutagenic and carcinogenic the conduction band and valence band of the semiconduc-
effects to the aquatic organisms (Perez et al. 2015). tor catalyst (TiO2). The property of semiconductor-free
MNZ is highly soluble in water (~ 10 mg mL−1 at 20 ° activated carbons to absorb photons and generate HO·
C) and can persist in the environment for several days radicals when exposed to UV irradiation in aqueous me-
(Wu and Fassihi 2005). The presence of MNZ in seven dium has been reported (Velasco et al. 2012, 2013;
major municipal wastewater treatment plant effluents Velo-Gala et al. 2015). However, the mechanism of radical
(2–40 μg L−1) has been reported (Gros et al. 2010), generation has not been explored in detail. On the other
which indicates the failure of conventional biological hand, the exposure of microorganisms in the UV-C/TiO2
treatment systems in stopping the entry of MNZ and and UV-C/GAC systems could facilitate the dimerization
similar other micropollutants along with the effluents. In of thymine (DoT) and the subsequent disruption of DNA
addition, the exposure of MNZ could lead to AMR replication process as shown in Eq. (6).
genes. Therefore, it is essential to design a suitable AMRg þ hv ð254 nmÞ→DoT→Cell death ð6Þ
treatment system to remove micropollutants including
MNZ and to control the propagation of AMR genes. The application of photocatalytic system for
The advanced oxidation processes (AOPs) such as real-time wastewater treatment depends on several fac-
ozonation, Fenton, photo-Fenton, photocatalysis, ultra- tors including physico-chemical characteristics of the
sound irradiation, electrochemical, and wet air oxidation micropollutant, energy and intensity of UV light, reac-
are suitable for the degradation of MNZ (Kanakaraju tion conditions (pH, pollutant concentration, catalyst
et al. 2014); however, photocatalysis employing UV in dosage, reusability of the catalyst, etc.), and ecotoxicity
the bandwidth of 200 to 280 nm (UV-C/TiO2 and UV-C/ of photoproducts formed (if any) in the treated water. In
GAC) is found to be very effective in MNZ and AMR the past, the efficacy of UV-TiO2 system in the removal
gene (AMRg) removals (Alexander et al. 2016; Guo of dyes, phenol, pesticides, solvents, and few pharma-
et al. 2013; Mckinney and Pruden 2012). ceutical compounds (Choquette-Labbe et al. 2014; Mar-
The photocatalytic system employing TiO2 catalyst tinez et al. 2011; Pereira et al. 2013; Verma et al. 2013;
has the following advantages: (1) TiO2 is chemically Zhu et al. 2012) under different experimental conditions
Water Air Soil Pollut (2017) 228:339 Page 3 of 12 339

were investigated. But the application of TiO 2 fraction, bulk density, specific surface area (SBET), and
photocatalysis, GAC as a catalyst, and their comparative total pore volume (V T ) of 2–5 mm, 0.4 kg m −3,
evaluation for antibiotics removal was not investigated 1010 m2 g−1, and 0.532 cm3 g−1, respectively, was
in detail. Moreover, photocatalytic degradation antibi- supplied by S.D. Fine Chemicals Limited, India.
otics using TiO2 and GAC as catalysts and the perfor-
mance of these systems based on the concept of electri-
cal energy per order (EEO) has not been explored in 2.2 Photolytic and Photocatalytic Experiments
detail. Therefore, the present study was focused (i) to
evaluate the performance of photocatalytic system using A double-walled cylindrical reactor with a working vol-
TiO2 and GAC (UV-C/TiO2 and UV-C/GAC) as cata- ume of 1.9 L (Fig. 1) was used for the photolytic and
lysts in the removal of MNZ; (ii) to distinguish the rates photocatalytic experiments. The UV lamps and probes
of photolytic, photocatalytic, and adsorption systems (pH, temperature, and oxidation-reduction potential
under different initial MNZ concentrations; (iii) to iden- (ORP)) were inserted into the reactor through the ports in
tify the degradation pathway of MNZ mediated by the lid of the reactor. An electronic overhead stirrer was
hydroxyl radicals; and (iv) to correlate the MNZ remov- mounted centrally and the reactor contents were mixed
al with energy consumption analysis in terms of EEO. continuously at a constant stirring speed. The size of GAC
used in the experiment was of 2–3 mm. The uniform
distribution of the GAC was achieved by the operation of
overhead stirrer at 500 rpm, whereas the uniform distribu-
2 Materials and Methods tion of TiO2 was achieved by operating at 200 rpm. The
light energy was supplied through two UV lamps each of
2.1 Chemicals 16 W power (λmax at 254 nm). The reactor was provided
with water circulation arrangement through the double
MNZ (99.9% purity) was obtained from Sigma-Aldrich, walls to control the increase in temperature of the reactor
India and used as received. Anatase form of TiO2 (size due to the supply of light energy. However, the experi-
less than 25 nm) having a surface area of 45–55 m2 g−1 ments in this investigation were conducted without circu-
and with a density of 3.9 g mL−1 at 25 °C was purchased lation of cooling water, i.e., uncontrolled temperature con-
from Sigma-Aldrich, India. Acetonitrile and deionized ditions. The experimental conditions of this investigation
water were used for HPLC analysis and they were are shown in Table 1. Photolytic experiments were con-
procured from Merck, Germany. GAC with particle size ducted at different initial MNZ concentrations, i.e., 0.1, 1,

Fig. 1 Schematic of batch


photocatalytic reactor
ELECTRONIC STIRRER

UV POWER PROBES
SUPPLY CONNECTED TO
pH/ORP/Temp
METER

DOUBLE WALLED
CYLINDRICAL
GLASS REACTOR
WITH WORKING
VOLUME OF 1.9 L
METRONIDAZOLE AND
TiO2 IN SUSPENSION
339 Page 4 of 12 Water Air Soil Pollut (2017) 228:339

Table 1 Experimental conditions of the study and their outcomes

Expt. MNZ Type of UV TiO2 GAC MNZ MNZ removal


runs concentration system power loading loading removal rate constant (×
(mg L−1) (W) (g/L) (g/L) at 120 min 10−2 min−1)
(%)

E-1 0.1 Photolysis 32 – – 39.2 0.39


E-2 1 32 – – 40.4 0.38
E-3 5 32 – – 38.4 0.33
E-4 15 32 – – 38.9 0.34
E-5 50 32 – – 24.5 0.23
E-6 100 32 – – 20.1 0.17
E-7 0.1 Photocatalysis 32 2.5 – 99.7 4.8
E-8 1 with TiO2 32 2.5 – 98.2 3.2
E-9 5 32 2.5 – 97.1 3.0
E-10 15 32 2.5 – 96.2 2.3
E-11 50 32 2.5 – 79.7 1.4
E-12 100 32 2.5 – 46.8 0.5
E-13 0.1 Photocatalysis 32 – 2.5 96.3 2.5
E-14 1 with GAC 32 – 2.5 94.7 2.3
E-15 5 32 – 2.5 89.7 1.9
E-16 15 32 – 2.5 77.6 1.28
E-17 50 32 – 2.5 72 1.3
E-18 100 32 – 2.5 36.3 0.4
E-19 0.1 Adsorption – – 2.5 23 0.29
E-20 1 – – 2.5 28 0.20
E-21 5 – – 2.5 27.4 0.22
E-22 15 – – 2.5 23.1 0.15
E-23 50 – – 2.5 18.3 0.17
E-24 100 – – 2.5 7 0.007

Note: the experiments were conducted in duplicate and the average values are reported

5, 15, 50, and 100 mg L−1, with continuous supply of working volume of 250 mL. The experiments were
32 W UV power (E-1 to E-6). The experiments were conducted at room temperature (~ 30 °C) with homog-
carried out for 120 min and the samples were collected at enous mixing in an orbital shaker for 120 min. The
5, 10, 15, 30, 60, and 120 min. The collected samples were aliquots were collected at regular time intervals and
analyzed for pH, temperature, ORP, total organic carbon analyzed for MNZ concentration. The outcome of the
(TOC), and MNZ concentration. The experiments were adsorption experiments were used to find out the rate of
repeated in the presence of photo catalysts each with the adsorption and MNZ adsorption capacity of GAC.
loading of 2.5 g L−1, i.e., TiO2 (E-7 to E-12) and GAC
(E-13 to E-18) with 32 W UV power. From all experi- 2.4 Analytical Methods
ments, the samples were collected at similar time intervals
and analyzed for various parameters. The pH, ORP, and temperature of the collected samples
were measured using bench-top meters (Oakton, India).
2.3 Adsorption Experiments The total organic carbon remaining in the samples was
analyzed by total organic carbon analyzer equipped with
The adsorption experiments (E-19 to E-24 as shown in non-destructive infrared analyzer (Shimadzu, Japan).
Table 1) were conducted using conical flasks at a The MNZ concentration was measured using
Water Air Soil Pollut (2017) 228:339 Page 5 of 12 339

high-performance liquid chromatography fitted with 3.2 MNZ Adsorption


C 18 220 mm (L) × 4.6 mm (I.D) silica-based
reverse-phase column and a UV-visible detector. The The experimental data corresponding to E-19 to
mobile phase used in the HPLC analysis was acetonitrile E-24 were used to calculate the equilibrium MNZ
and water (60:40), which was pumped at a rate of adsorption concentration on GAC using Eq. (9).
1 mL min−1. Where Ce, V, and W represent equilibrium MNZ
The photocatalytic degradation products of MNZ in concentration at equilibrium time (te), reaction vol-
the samples (E-7 to E-12) were identified using a liquid ume, and weight of GAC. Moreover, the data was
chromatography fitted with ion-trap mass spectrometer used to estimate the maximum MNZ adsorption
(LC-MS) (Thermoscientific, USA) and operated under capacity of GAC using Langmuir (Eq. 10),
positively induced ion scanning-mode. The analyzer Freundlich (Eq. 11), and Sips (Eq. 12) isotherms.
was maintained at an electro spray voltage of +1.5 kV
V
ion source with a capillary temperature of 350 °C. The qe ¼ ðC o −C e Þ ð9Þ
W
samples were injected manually and the obtained spec-
trums were used to identify the degradation products.
The TiO2 and GAC samples (fresh and samples qmL K L C e
collected from the reactor after experimentation) were qe¼ ð10Þ
1 þ K LCe
dried and coated with gold. Subsequently, the surface
morphology of samples was visualized by field emis-
sion scanning electron microscopy (SEM) at an accel- .
1 nL
erating voltage of 30 kV (Genisis 2000, Emcrafts). qe¼ K F C e ð11Þ
Moreover, the elemental composition of the samples
was quantitatively determined using energy dispersive
spectroscopy (EDS, Genesis 2000). qms : K S C m s
e
qe¼ ð12Þ
1 þ K sCe ms

In Eq. (10), qmL is the amount of adsorption


3 Data Interpretation corresponding to the complete monolayer coverage
and KL is the Langmuir isotherm constant. In Eq.
3.1 Rate of MNZ Removal (11), 1/nL and KF are the adsorption intensity and
Freundlich isotherm constant, respectively. In Sips
The photocatalytic removal efficiency of MNZ in the isotherm, qms is maximum adsorption capacity, Ks is
system was calculated as shown in Eq. (7). the sips isotherm constant, and ms is the model
exponent. At lower sorbate concentrations, Eq. (12)
C 0 −C t can reduce to Freundlich isotherm while at higher
η¼  100 ð7Þ
C0 concentrations the sips model can predict the mono-
layer sorption capacity as like the Langmuir iso-
Where C0 and Ct represent the initial MNZ concentra- therm. The rate of adsorption, indicating the resi-
tion and final MNZ concentration at any time (t), respec- dence time of the compound at the solid solution
tively. The relationship between degradation rate and the interface was found out by using kinetic models. In
initial MNZ concentration (C0) in the photocatalytic pro- order to find out the role of GAC on MNZ sorption,
cess can be described by Langmuir-Hinshelwood (L-H) Lagergren’s pseudo-first order (Eq. 13) and
model (Hapeshi et al. 2010; Prados-Joya et al. 2011; pseudo-second order model (Eq. 14) were applied.
Safari et al. 2015; Vishnuganth et al. 2016). The simpli- Moreover, Weber Morris Eq. (15) was used to iden-
fied form of L-H model as shown in Eq. (8) was used to tify the role of intra-particle diffusion in MNZ ad-
find out the rate of MNZ removal (k1). sorption. The rate and order of the adsorption reac-
tion was further validated by Eq. (16) where dC and
Ct CA are the difference and average MNZ concentra-
ln ¼ −k 1 t ð8Þ
C0 tions measured at the successive time intervals. The
339 Page 6 of 12 Water Air Soil Pollut (2017) 228:339

value of n denotes the order of the reaction kinetic model (Eq. 8) has well fitted the experimental data
(Table 2). with high correlation coefficient value (r2 > 0.98). The rate
constant (k1) was determined from the slope of the linear-
k1
log ðqe −qt Þ ¼ logqe − t ð13Þ ized first-order plot of -ln (C0/C) versus time, and the
2:303
values are given in Table 1. The observation of relatively
higher k1 values in E-7 to E-12 compared to E-1 to E-6 is
t 1 1 an indication of accelerated MNZ degradation by the
¼ þ t ð14Þ HO·-mediated oxidation reaction. However, a decline in
qt K 2 q2e qe
k1 value with increase in MNZ concentration can be
noticed in Table 1, which could be due to the availability
. of fewer quantity of HO· radicals per mole of MNZ at
1 2
qt ¼ K id t þc ð15Þ higher initial MNZ concentrations. Moreover, the degra-
dation intermediates of MNZ might have accumulated on
the surface of the catalyst causing a negative effect in the
  utilization of OH· radicals or positive holes in the valence
−dc
log ¼ n logC A þ logk 1 ð16Þ band of TiO2 surface (Cai and Hu 2017; Farzadkia et al.
dt
2015). The increase in the value of removal rate constant is
several times higher in both photolytic (2.5 times) and
photocatalytic (9.6 times) systems at lower concentration
(0.1 mg L−1) compared to higher MNZ concentration
4 Results and Discussion (100 mg L−1). A marginal decline in pH was noticed in
the photolytic and photocatalytic systems during the first
4.1 MNZ Removal by Photolysis and Photocatalysis 20 min of reaction; however, the pH trends of both systems
were found to be similar (Fig. 2). The decline in pH was by
The outcome of photolysis (E-1 to E-6) and photocatalysis the dissolution of carbon-dioxide (CO2) produced as a
(E-7 to E-18) experiments are shown in Table 1. A max- result of partial mineralization, which was evidenced by
imum of 40.4% MNZ removal was observed in the pho- the LC-MS analysis. On the other hand, MNZ removal
tolysis experiments corresponds to E-2 (initial MNZ con- might not be significant at other pH conditions due to the
centration was at 1 mg L−1). A decline in MNZ removal nature of MNZ and TiO2 at different solution pHs. The
was observed while increasing the MNZ concentration pHzpc of TiO2 is 6.52 and pKa of MNZ is 2.55. At basic
beyond 1 mg L−1. The ORP profiles corresponding to pH, both TiO2 and MNZ possess negative charge and at
selective photolytic and photocatalytic experiments are acidic pH both are positively charged. Therefore, at neutral
shown in Fig. 2. It can be observed from the figure that pH, no repulsive forces might be developed and as a result,
ORP values are significantly higher in photocatalytic sys- maximum degradation could be achieved (Farzadkia et al.
tems (171–323 mV) compared to photolytic systems 2015).
(174–241 mV). This could be due to greater quantity of The applicability of GAC as a photocatalyst was ex-
hydroxyl radicals (HO·) generated in the photocatalytic amined in E-13 to E-18 and the outcomes are shown in
system. The MNZ removal efficiencies observed in E-7 Table 1. A maximum of 96.3% removal was observed in
and E-8 were 1.5 times higher than that observed in E-2. E-13, whereas 36.3% was observed at the highest initial
As observed in photolytic systems, similar decline in MNZ MNZ concentration adopted for UV-C/GAC experi-
removal efficiency was observed beyond 0.1 mg L−1. ments. Similar ORP and pH trends were observed in
However, the lowest MNZ removal observed in the pho- UV-C/GAC system as like UV-C/TiO2 (Fig. 2). The
tocatalytic systems (E-12) was relatively higher than the removal efficiency of GAC photocatalytic system was
maximum MNZ removal observed in photolytic system 146% higher compared to the photolytic system; howev-
(E-2). This could be attributed to MNZ oxidation under er, it was slightly lesser (3.5%) compared to photocata-
irradiation or hydrolysis in the photolytic system, whereas lytic system with lowest initial MNZ concentration, i.e.,
MNZ oxidation was mediated by HO· in photocatalytic E-7. Similar trend was observed under other MNZ con-
system. As a result, higher MNZ removal rates were centrations, which show that GAC can be used as a
observed in photocatalytic system. The pseudo-first-order photocatalyst apart from it being used as an adsorbent
Water Air Soil Pollut (2017) 228:339 Page 7 of 12 339

Table 2 Analysis of MNZ adsorption on GAC

Experimental runs Adsorption rate constant, k1 (min−1) based on Outcomes of adsorption isotherm modeling

Eq. (13) Eq. (16) Type of isotherm Parameters

E-19 – 2.09 (1.4) Freundlich KF = 1.145 mg g−1


nL = 0.127
E-20 0.31 1.93 (1.7) r2 = 0.947
E-21 0.05 0.53 (1.3) Langmuir qmL = 3.165 mg g−1
KL = 0.067 L mg−1
E-22 0.009 0.60 (1.2) r2 = 0.904
E-23 0.02 0.59 (1.1) Sips Ks = 2.16 L mg−1
ms = 0.647
E-24 0.06 3.84 (0.7) r2 = 0.962

Note: the values in parenthesis indicate the order of the reaction as calculated from Eq. (16)

(Velasco et al. 2012). The value of k1 was 2.0 min−1 in observed in E-20 and E-21, which was much lesser com-
E-13 but it was reduced by five times at the highest MNZ pared to the MNZ removal observed under photolysis and
concentration investigated (E-18). photocatalysis systems. The value of MNZ removal rate
constant (k1) in the GAC adsorption was also calculated
4.2 MNZ Removal by Adsorption using the first-order reaction (Eq. 8). As expected, the k1
values were found to lower compared to the other systems
The SEM and EDS analysis confirmed the presence of (E-1 to E-18). In addition, the rate of adsorption was
MNZ on the surface pores of GAC (Fig. 3). Owing to the determined using Eqs. (13)–(16) and it was found that
neutrality of MNZ (pKa = 2.5) in the pH range of 6 to 7, the the first-order equation was found to be suitable compared
adsorption could be mainly by physical forces/phenome- to other kinetic models (data not shown). The calculated
non. Based on the profiles of residual MNZ concentration adsorption rate constant values based on Eqs. (13) and (16)
in the reaction mixtures, the maximum MNZ adsorption and the order of reaction (n) based on Eq. (16) are shown in
capacity of GAC were found out and the values are given Table 2. The value of n indicates that the adsorption
in Table 2. A maximum MNZ removal of around 28% was depends on MNZ concentration at lower MNZ

Fig. 2 pH/ORP profiles for UV- 9 350


C only, UV-C/TiO2, and UV-C/
8
GAC experiments for degradation
of 15 mg L−1 of MNZ 7
300

6
250
ORP (mV)

5
pH

4
200
3

2 pH trend - E-4 pH trend - E-10


150
pH trend - E-16 ORP trend - E-4
1
ORP trend - E-10 ORP trend - E-16
0 100
0 20 40 60 80 100 120
Time (min)
339 Page 8 of 12 Water Air Soil Pollut (2017) 228:339

Fig. 3 SEM images of GAC (a) before adsorption and (b) after adsorption, and EDS images of GAC (c) before adsorption and (d) adsorption

concentration, whereas at the highest MNZ concentration converge to Freundlich’s isotherm for the range of
(adopted in this investigation), the reaction order does not MNZ concentration adopted in this study.
depend on initial adsorbate concentration.
The Freundlich isotherm was found to be suitable to 4.3 Comparison of MNZ Removal Based on Energy
represent the MNZ adsorption on GAC. A smaller value Consumption Analysis
of 1/nL indicates the formation of strong bond between
the adsorbate and adsorbent. The maximum MNZ ad- The potential of each process, i.e., photolysis (E-1 to E-6),
sorption capacity of GAC was calculated as photocatalysis with TiO2 and GAC (E-7 to E-18), and
1.145 mg g−1 based on E-19 to E-24. It is envisaged adsorption on GAC (E-19 to E-24), in terms of MNZ
that the adsorption sites of GAC might be occupied at removal rates is shown in Fig. 4a. Based on the figure, it
much faster rate at higher MNZ concentration (E-24); as can be concluded that UV-C/TiO2 and UV-C/GAC are
a result, the overall adsorption could not be represented highly efficient processes with greater rate constant values
by the first-order equation. Moreover, the adsorption compared to photolysis (UV-C only) and adsorption on
might have followed multi-layer rather than monolayer GAC. Moreover, it can be noticed that the removal rates
as expressed by Langmuir’s isotherm at higher concen- were fairly equal for UV-C/TiO2 and UV-C/GAC systems
trations. On the other hand, the smaller ms value calcu- at higher MNZ concentrations (50 and 100 mg L−1). How-
lated based on Sips isotherm indicate that it can ever, energy consumption and EEO analysis are better
Water Air Soil Pollut (2017) 228:339 Page 9 of 12 339

Fig. 4 a Removal rates and b


EEO values for various systems
a
0.06
under different initial MNZ

Rate of MNZ removal (min-1)


concentrations 0.05 Photolysis
Photocatalysis with TiO2
0.04 Photocatalysis with GAC
Adsorption
0.03

0.02

0.01

0
0.1 1 5 15 50 100
-1
MNZ concentration (mg L )

b
1.2

Photolysis
1
Photocatalysis with TiO2
EEO (KWh m-3 order-1)

Photocatalysis with GAC


0.8 Adsorption

0.6

0.4

0.2

0
0.1 1 5 15 50 100
MNZ concentration (mg L-1)

suitable in the cost evaluation of different processes. The order−1) (Fig. 4b). The EEO value was found to be increas-
EEO of all experimental runs were calculated based on Eq. ing with the increase in MNZ concentrations; however, no
(17) and the values were used to identify the effectiveness significant difference in EEO value was observed between
of the system (Vishnuganth et al. 2016). 0.1 and 15 mg L−1. This shows that photocatalytic system
with TiO2 can be used for the removal of micropollutants
P  t  1000
E EO ¼ ð17Þ from both industrial wastewater and contaminated surface/
C0
V  60  log ground waters. However, the effect of other ions and solids
Ct in the removal of micropollutants in photocatalytic system
need to be analyzed.
Where P is the power consumed by the system
(expressed in Watt) and V is the volume of the reactor 4.4 MNZ Degradation Pathway
(i.e., 1.9 L). The term P is the major component, which
accounts for the operational cost of the photolytic and The LC-MS spectrums for the samples (E-10) were
photocatalytic systems for a given time in the removal of analyzed at 0 and 120 min as shown in Fig. 5a. Based
the target micropollutant, i.e., MNZ. From the EEO values, on LC-MS analysis, a plausible reaction sequence of
it was evident that photocatalytic systems with TiO2 and MNZ mineralization under TiO2 slurry photocatalytic
GAC were highly energy efficient (0.03 to 0.34 kWh m−3 system was proposed as shown in Fig. 5b. Previously, it
order−1) compared to photolysis (0.15 to 0.4 kWh m−3 was reported that the first step in the degradation of
order−1) and adsorption systems (0.44 to 0.99 kWh m−3 MNZ was mainly by hydroxylation due to HO· radical
339 Page 10 of 12 Water Air Soil Pollut (2017) 228:339

a b
D:\Raw Data\...\met 0 min_160212230904 3/4/2016 7:53:16 PM met 0 min D:\Raw Data\...\met 120 min_160212230904 3/4/2016 7:55:08 PM met 120 min

met 0 min_160212230904 #98 RT: 0.92 AV: 1 NL: 7.29E6 met 120 min_160212230904 #259 RT: 2.39 AV: 1 NL: 1.45E7
T: FTMS + c ESI Full ms [100.00-2000.00] T: FTMS + c ESI Full ms [100.00-2000.00]
172.0719 116.9856
100 100

95 95

90 90

85 85

80 80

75 75

70 70

65 65

60 60

Relative Abundance
Relative Abundance

55 55

50 50

45 45

40 40

35 35 149.0120
30 30

25 103.9781 25
132.9581
20 20 147.0553
116.9860 103.9779 129.0301 134.9964
15 130.5314 15 113.9765 118.9788
136.0045 155.0294
146.5446 123.0031
10 121.5261 10
153.5254 170.0326 158.0124
105.5129 128.0456 198.0190 164.9845
5 161.0124 175.1193 5
112.9576 191.0455 109.0238 141.9585 174.9914 187.0013 198.9849
0 0
100 110 120 130 140 150 160 170 180 190 200 100 110 120 130 140 150 160 170 180 190 200
m/z m/z

c N Cleavage of the N OH
N -NO3
O ethanol group
+
H3C
N
N H3C OH H3C OH
-
N N
O H (C)
Breaking of
HO
OH HO imidazole moiety -NO3

(A) (B)

O
NH2
O
(D)
OH

CO2
O O
+
Simpler OH
NO3- / NO2- +
aliphatics H2C
+ H2N OH
H2O
(F) (E)
Fig. 5 a LC-MS spectrum of sample collected from E-10 at 0 min, b LC-MS spectrum of sample collected from E-10 at 120 min, and c
proposed metronidazole mineralization pathway in photocatalytic system

produced in the photocatalytic system and subsequent The other major degradation intermediates identified
formation of compound-B (Fig. 5b). Subsequently, the from LC-MS spectrums were found to have m/z values
cleavage of lateral N-ethanol group has resulted in the 116 and 149. The formation of heterocyclic
formation of compound-C as proposed by Perez et al. compound-D (3-carbomoyl prop 2-enoic acid, m/z
(2015). However, both compounds-B and -C 116) was due to the attack by OH· which led to the
(3-(2-hydroxy-ethyl)-2-methyl-3H-imidazol-4-ol and breaking of imidazole moiety and subsequent, release of
2-methyl-3H-imidazol-4-ol, respectively) were not sta- NO3− ion. Further, the removal of lateral groups (i.e., –
ble with time (i.e., no peaks were observed at 120 min). CH = CH–) from compound-D (i.e., compound-E) and
Water Air Soil Pollut (2017) 228:339 Page 11 of 12 339

the addition of compound-F (2-propen-1-ol, i.e., formed 2.511 for TiO2 catalyst after 15 min of exposure. The
by addition of OH to the lateral groups) has shown m/z raw sample from E-10 (0 min), treated sample from
value of 149. The compound-E is a short-linear aliphatic E-10 after 120 min (without settling), and treated
carboxylic acid produced by the oxidation of samples from E-10 after 6 and 12 h settling have shown
compound-D. This particular observation correlates an inhibition effect of 18.1, 13.8, 11.9, and 9.52%,
well with the decline in pH of the system toward acidic respectively, at 15-min exposure time. The difference
range (from pH 8 to pH 5 after 120 min). in percentage inhibition between 0- and 120-min sam-
The sample collected from E-10 was analyzed for ples of E-10 was due to the presence of MNZ metabo-
TOC and the presence of nitrate/nitrite ions using ion lites. But the toxic levels of samples are lesser than the
chromatography (Dionex, USA). The results show that potential toxic limits. However, the long-term persis-
considerable quantity of carbon molecules present in tence of MNZ and their metabolites can lead to bacterial
MNZ was removed from the system as CO2 gas as a resistance, which needs further attention.
result of photocatalytic oxidation of MNZ. The total
carbon removal in the system was 66.4%, which indi- 6 Conclusions
cate a partial mineralization of MNZ and formation of
MNZ by-products in the system. On the other hand, ion The application of TiO2 and GAC as catalysts in the UV-
chromatography analysis confirmed the formation of C photocatalytic system has been investigated. The
nitrate in the system. Therefore, the effluent from the MNZ removal efficiencies in UV-C/TiO2 and UV-C/
photocatalytic system needs to be treated for nitrate or GAC systems were fairly equal at lower MNZ concen-
the stability of nitrate in the same system needs to be trations (0.1–5 mg L−1). However, decline in MNZ
investigated. The experiments and correlations corre- removal efficiency was observed in both systems be-
sponding to the formation of nitrate ion in the system yond an initial MNZ concentration of 15 mg L−1. The
and the possible theoretical nitrogen release in the sys- MNZ removal was modeled using the first-order rate
tem are in progress. equation and the efficiency of various systems were
analyzed by EEO values. The EEO values were in good
correlation with the removal rate constants. The MNZ
mineralization pathway in the photocatalytic system
5 Microtoxicity Studies was proposed based on LC-MS analysis. The release
of ionic compounds of nitrogen and degradation inter-
The bioluminescence inhibition assay was conducted mediates of MNZ would be of concern in the effluent.
using a marine bacterium Vibrio fischeri to assess the As a whole, it can be concluded that the designed
toxicity of MNZ, TiO2, and the treated samples (after photocatalytic system is highly efficient in the removal
settling TiO2 catalyst for 6 and 12 h) from selected of micropollutants from industrial wastewater and con-
experimental runs. The bacterial luminescence readings taminated surface/ground waters.
were measured using a Microtox model 500 analyzer
(Modern Water Inc., UK) at different exposure time. Acknowledgements The authors acknowledge the financial
The percentage inhibition effect of the samples to the support from Centre for Industrial Consultancy and Sponsored
Research (ICSR), IIT Madras to execute this research work (Grant
bacterium is calculated as toxicity units (TU) using No: CIE/14-15/832/NFIG/SMAT and CIE/14-15/650/NFSC/
Eq. (18). The 81.9% basic protocol (consisting of four SMAT).
test dilutions) was carried out to evaluate the ecotoxicity
of the aqueous suspensions. The possible interference of
TiO2 in the luminescence measurements if any were
accounted by measuring the absorbance of the tested References
suspension without bacterium addition.
. Alexander, J., Karaolia, P., Fatta-Kassinos, D., & Schwartz, T.
Toxicity unit ðTU50 Þ ¼ 100 EC50 ð18Þ (2016). Impacts of advanced oxidation processes on
microbiomes during wastewater treatment. In D. Fatta-
Kassinos, D. D. Dionysiou, & K. Klaus (Eds.), Advanced
Accordingly, the toxicity unit of MNZ was found to treatment technologies for urban wastewater reuse (pp. 129–
be 1.573 after 60-min exposure time, whereas it was 144). Switzerland: Springer.
339 Page 12 of 12 Water Air Soil Pollut (2017) 228:339

Andreozzi, R., Caprio, V., Insola, A., & Marotta, R. (1999). Lupo, A., Coyne, S., & Berendonk, T. U. (2012). Origin and
Advanced oxidation processes (AOP) for water purification evolution of antibiotic resistance: The common mechanisms
and recovery. Catalysis Today, 53(1), 51–59. of emergence and spread in water bodies. Frontiers in
Asha, C. R., & Kumar, M. (2015). Sulfamethoxazole in poultry Microbiology, 3, 18.
wastewater: identification, treatability and degradation path- Martinez, C., Canle, M., Fernandez, M. I., Santaballa, J. A., &
way determination in a membrane-photocatalytic slurry reac- Faria, J. (2011). Kinetics and mechanism of aqueous degra-
tor. Journal of Environmental Science and Health Part-A dation of carbamazepine by heterogeneous photocatalysis
Toxic/Hazardous Substances & Environmental Engineering, using nanocrystalline TiO2, ZnO and multi-walled carbon
50(10), 1011–1019. nanotubes–anatase composites. Applied Catalysis B:
Bendesky, A., Menendez, D., & Ostrosky-Wegman, P. (2002). Is Environmental, 102, 563–571.
metronidazole carcinogenic. Mutation Research, 511(2), Mckinney, W. C., & Pruden, A. (2012). Ultraviolet disinfection of
133–144. antibiotic resistant bacteria and their antibiotic resistant genes
Cai, Q., & Hu, J. (2017). Decomposition of sulfamethoxazole and in water and wastewater. Environmental Science &
trimethoprim by continuous UVA/LED/TiO2 photocatalysis: Technology, 46(24), 13393–13400.
decomposition pathways, residual antibacterial activity and Pereira, L., Pereira, R., Oliveira, C. S., Apostol, L., Gavrelescu,
toxicity. Journal of Hazardous Materials, 323, 527–536. M., Pons, M. N., et al. (2013). UV/TiO2 photocatalytic deg-
Choquette-Labbe, M., Shewa, A. W., Lalman, A. J., & radation of xanthine dyes. Photochemistry and Photobiology,
Shanmugam, R. S. (2014). Photocatalytic degradation of 89(1), 33–39.
phenol and phenol derivatives using a nano-TiO2 catalyst: Perez, T., Garcia-Segura, S., El-Ghenymy, A., Nava, J. L., & Brillas,
Integrating quantitative and qualitative factors using response E. (2015). Solar photoelectro-Fenton degradation of the metro-
surface methodology. Water, 6(6), 1785–1806. nidazole using a flow plant with a Pt/air diffusion cell and a CPC
Coutu, S., Wyrsch, V., Wynn, H. K., Rossi, L., & Barry, D. A. photoreactor. Electrochimica Acta, 165, 173–181.
(2013). Temporal dynamics of antibiotics in wastewater treat- Petrovic, M., Radjenovic, J., & Barcelo, D. (2011). Advanced
ment plant influent. Science of the Total Environonment, 458- oxidation processes (AOPs) applied for wastewater and
460, 20–26. drinking water treatment, elimination of pharmaceuticals.
Holistic approach to Environment, 12, 63–74.
Farzadkia, M., Bazrafshan, E., Esrafili, A., Yang, J. K., & Shirzad-
Prados-Joya, G., Sanchez-Polo, M., Rivera-Utrilla, J., & Ferro-
Siboni, M. (2015). Photocatalytic degradation of metronida-
Garcia, M. (2011). Photodegradation of nitroimidazoles in
zole with illuminated TiO 2 nanoparticles. Journal of
aqueous solution by ultraviolet radiation. Water Research,
Environmental Health Science & Engineering, 13, 35.
45(1), 393–403.
Gros, M., Petrovic, M., Ginebreda, A., & Barcelo, D. (2010).
Safari, G. H., Hoseini, M., Seyedsalehi, M., Kamani, H., Jaafari,
Removal of pharmaceuticals during wastewater treatment
J., & Mahvi, A. H. (2015). Photocatalytic degradation of
and environmental risk assessment using hazard indexes.
tetracycline using nanoized titanium dioxide in aqueous so-
Environment International, 36(1), 15–26.
lution. International Journal of Environmental Science &
Guo, M., Yuan, Q., & Yang, J. (2013). Microbial selectivity of UV Technology, 12(2), 603–616.
treatment on antibiotic-resistant heterotrophic bacteria in sec- Velasco, F. L., Fonseco, M. I., Parra, B. J., Lima, C. J., & Ania, O.
ondary effluents of wastewater treatment plants. Water C. (2012). Photochemical behavior of activated carbons un-
Research, 47(16), 6388–6394. der UV irradiation. Carbon, 50(1), 249–258.
Guo, J., Selby, K., & Boxall, A. B. (2016). Effects of antibiotics on Velasco, F. L., Maurino, V., Laurenti, E., Fonseca, M. I., Lima, C.
the growth and physiology of chlorophytes, cyanobacteria, J., et al. (2013). Photoinduced reactions occurring on activat-
and a diatom. Archives of Environmental Contamination and ed carbons. A combined photooxidation and ESR study.
Toxicology, 71(4), 589–602. Applied catalysis A: General, 15, 1–8.
Hapeshi, E., Achilleos, A., Vasquez, M. I., Michael, C., Velo-Gala, I., Lopez-Penalver, J. J., Sanchez-Polo, M., & Rivera-
Xekoukoulotakis, N. P., Mantzavinos, D., et al. (2010). Utrilla, J. (2015). Role of activated carbon on micropollutants
Drugs degrading photocatalytically: Kinetics and mecha- degradation by different radiation processes. Mediterranean
nisms of ofloxacin and atenolol removal on titania suspen- Journal of Chemistry, 4, 68–80.
sions. Water Research, 44(6), 1737–1746. Verma, A., Sheoran, M., & Toor, A. P. (2013). Titanium dioxide
Kanakaraju, D., Glass, B. D., & Oelgemoller, M. (2014). Titanium mediated photocatalytic degradation of malathion in aqueous
dioxide photocatalysis for pharmaceutical wastewater treat- phase. Indian Journal of Chemical Technology, 20, 46–51.
ment. Environmental Chemistry Letters, 12(1), 27–47. Vishnuganth, M. A., Remya, N., Kumar, M., & Selvaraju, N.
Kimosop, S. J., Getenga, Z. M., Orata, F., Okello, V. A., & (2016). Photocatalytic degradation of Carbofuran by TiO2-
Cheruiyot, J. K. (2016). Residue levels and discharge loads coated activated carbon: Model for kinetic, energy per order
of antibiotics in wastewater treatment plants (WWTPs), hos- and economic analysis. Journal of Environmental
pital lagoons and rivers within Lake Victoria Basin, Kenya. Management, 181, 201–207.
Environmental Monitoring and Assessment, 188(9), 532. Wu, Y., & Fassihi, R. (2005). Stability of metronidazole, tetracy-
Lien, L. T., Hoa, N. Q., Chuc, N. T., Thoa, N. T., Phuc, H. D., cline HCl and famotidine alone and in combination.
Diwan, V., et al. (2016). Antibiotics in wastewater of a rural International Journal of Pharmaceutics, 290, 1–13.
and an urban hospital before and after wastewater treatment, Zhu, X., Zhou, D., Cang, L., & Wang, Y. (2012). TiO2 photocat-
and the relationship with antibiotic use—a one year study alytic degradation of 4-chlorobiphenyl as affected by solvents
from Vietnam. International Journal of Environmental and surfactants. Journal of Soils and Sediments, 12(3), 376–
Research and Public Health, 13(6), 588. 385.

You might also like