Comprehensive Treatment of Urban Wastewa

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Environmental Management 266 (2020) 110469

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Research article

Comprehensive treatment of urban wastewaters using electrochemical


advanced oxidation process
Bokam Rajasekhar a, Urmika Venkateshwaran b, Nivetha Durairaj b, Govindaraj Divyapriya a,
Indumathi M. Nambi a, *, Angel Joseph a
a
Environmental and Water Resources Engineering Division, Indian Institute of Technology Madras, Chennai, 600036, India
b
Easwari Engineering College, Chennai, 600089, India

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

Keywords: This study mainly focuses on the efficiency of anodic oxidation process (Ti/Sb-SnO2/PbO2 as anode and stainless
Ti/Sb-SnO2/PbO2anode steel as the cathode) in treating two different streams of urban wastewater, one from the influent of sequence
Electrochemical oxidation batch reactor (WW1) and other from the effluent of constructed wetland (WW2). The effect of different opera-
Urban wastewater
tional parameters such as current density, hydraulic retention time, exposed electrode surface area, phosphorous,
Removal of micropollutants
Removal of nutrients
ammonia-nitrogen, nitrates, and coliform bacteria was studied. For an optimized current density of 30 mA/cm2
Electrochemical disinfection and an electrode surface area of 30 cm2, almost complete removal of COD and ammonia-nitrogen were achieved
with both wastewaters (WW1 & WW2), while in case of phosphorous, 50% and 98% removal efficiencies were
observed. Electrode deposition was analyzed using SEM-EDS and XRD, which confirms the presence of calcium
and magnesium phosphates on the surface on the anode, which attributes to the phosphate removal. Electro-
chemical disinfection studies showed that complete inactivation of bacteria takes place within 30 min for WW1
and 60 min for WW2, and the cell morphological changes were studied using SEM analysis. Degradation of
different micropollutants present in the wastewaters was evaluated with the aid of GC-MS. ICP - MS analysis
confirmed that there was no leaching of lead from the anode surface, and the lead which is already present in the
wastewater gets reduced to a permissible level, which further increases the treatment efficiency. Hence cleaner
and comprehensive treatment of real urban wastewaters through anodic oxidation process was successfully
demonstrated in this work.

1. Introduction the electrochemical oxidation (EO) treatment process, which comes


under electrochemical advanced oxidation processes (EAOP). EO has
Pollution of surface and groundwater with organic matter, phos- advantages such as requirement of no chemicals for the treatment, no
phorus, nitrates, ammonia-nitrogen, heavy metals, and emerging pol- sludge production, comparatively lower investment cost and high
lutants (pharmaceuticals, pesticides and personal care products) is a pollutant removal efficiency, ease of the process, operation at ambient
growing concern globally. The conventional treatment processes (bio- conditions and low space requirements (Deng et al., 2019; Gargouri
logical and physio-chemical treatment processes) which include aerobic et al., 2014; Panizza and Cerisola, 2009).
and anaerobic treatment, coagulation and flocculation, adsorption and The mechanism of contaminant degradation in the EO involves
ozonation have limited applications in the case of refractory (chemical mainly two steps: (i) in-situ generation of hydroxyl radicals (�OH) at
and biological stability) organic pollutants (Liu et al., 2019; Rosal et al., electrode surface (Eq. (1)); (ii) oxidation of pollutants by the action of
2010). Though these organic contaminants are present in low concen- �OH radicals (Eq. (2)) (Martinez-Huitle and Ferro, 2006).

trations, they are chronic and hazardous owing to their toxicity, carci-
nogenicity, and mutagenicity (Garcia-Segura et al., 2018a). Hence M þ H2O → M(�OH) þ Hþ þ e (1)
alternative and efficient treatment technologies for the removal of these M(�OH) þ R → M þ mCO2 þ nH2O þ H þ e þ
(2)
persistent organic pollutants (POP’s) and reuse of water resources are
required (Garcia-Segura et al., 2018a). One such emerging technology is The above reaction is accompanied by a competitive reaction of the

* Corresponding author.
E-mail address: indunambi@iitm.ac.in (I.M. Nambi).

https://doi.org/10.1016/j.jenvman.2020.110469
Received 10 November 2019; Received in revised form 11 March 2020; Accepted 20 March 2020
Available online 17 April 2020
0301-4797/© 2020 Elsevier Ltd. All rights reserved.
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

conversion of hydroxyl radicals to oxygen without the participation of Pvt. Ltd., Mumbai, India) was used as a medium for the growth of mi-
the anode by Eq. (3). croorganisms. AquaSnap™ ATP Water Testing Device (Weber Scientific)
was used to measure ATP in the solution. HygienaAquaSnap™ Total
M(�OH) þ H2O → M þ O2 þ 3Hþ þ 3e (3)
measures both free ATP in solution and microbial ATP. AquaSnap Free
Hydroxyl radical, having a standard redox potential of 2.8 V, is the analyses ATP that is free in solution. Solid-phase extraction (SPE) car-
main feature of advanced oxidation processes as it reacts non-selectively tridges (Oasis HLB 6 cc) were procured from Waters Corporation, Mil-
with a wide group of organic and inorganic contaminants (Liu et al., ford, Massachusetts, USA.
2019). Wastewater samples were collected from two different locations in
It has been reported that anodes with low oxygen overpotential Chennai, Tamil Nadu, India between December 2018 and March 2019.
(active electrodes), such as IrO2, RuO2, and Pt oxidize the organic con- Wastewater sample 1 (WW1) was collected from the sequencing batch
taminants partially and hence less effective (Panizza and Cerisola, reactor (SBR) tank influent (before biological treatment) of a 4 MLD
2009). On the other hand, complete elimination of organic pollutants sewage treatment plant at IIT Madras, Chennai, India. The nature of
with optimum current efficiency can be firmly obtained by employing WW1 was a mixture of grey and black waters discharged from student
the high oxygen over potential anodes (non-active electrodes) such as hostels and was having high loads of COD, nutrients, and coliform
Boron doped diamond, BDD (Zhu et al., 2008; Urtiaga et al., 2014), SnO2 bacteria (Table 1). Wastewater sample 2 (WW2) was collected from a
(Vargas et al., 2014) and PbO2 (Dai et al., 2014). BDD, although effective different location, which is a constructed wetland site at the Chennai
in mineralization of pollutants, is an expensive anode due to an Mathematical Institute, SIPCOT, Chennai, India. Wastewater from the
energy-intensive process of its fabrication, whereas the lifetime of Institute hostels was collected in an Inlet chamber and passed through a
Sb-SnO2 is less. In the non-active anodes, PbO2 electrodes have a greater series of backfill reactors to the constructed wetland, which consists of
potential for electrochemical treatment applications. This is because of Canna plants. The effluent from this wetland was collected in storage
their low cost, easy fabrication, long life and their ability to breakdown tanks from which the sample was withdrawn (WW2). Wastewater
recalcitrant pollutants without compromise in mineralization efficiency samples were frequently collected and refrigerated at 4 � C until analysis
(Niu et al., 2012; Xia and Dai, 2018). and experiments.
In literatures, PbO2 anodes have been used to treat the diversified
organic pollutants such as pharmaceuticals (Bian et al., 2019; Dai et al., 2.2. Instrumentation and analysis
2014), dyes (Elaissaoui et al., 2019; Xia et al., 2020), phenolic com-
pounds and other emerging pollutants (Niu et al., 2012; Xia and Dai, Chemical oxygen demand was analyzed using closed reflux, titri-
2018; Rao and Venkatarangaiah, 2014). However, most of these studies metric method (Method name: 5220 CHEMICAL OXYGEN DEMAND)
utilized the PbO2 anodes to treat the synthetic wastewaters involving (APHA, 2017). COD digestion was done in HACH DRB200: digital
single pollutant systems. The studies on evaluating the potential of PbO2 reactor block digestor. HACH HQD portable meter was used to measure
anode to treat the real wastewaters were found to be scant. The few the pH and electrical conductivity of wastewater. Stannous Chloride
studies, which used the PbO2 electrode for real wastewaters include method was used for phosphorous analysis, using UV-VIS Spectropho-
landfill leachates (Chiang et al., 1995; Fernandes et al., 2014, 2016; tometer at 690 nm (Method name: 4500-P PHOSPHORUS) (APHA,
Mandal et al., 2020), coking wastewater (Ma et al., 2012), reactive 2017). Phenate Method was used for ammonia-nitrogen analysis, using
yellow dye in a real wastewater (Gaber et al., 2013), oilfield produced UV-VIS Spectrophotometer at 640 nm (Method name: APHA 4500-NH3
waters (Gargouri et al., 2014) and car wash wastewater (Panizza and NITROGEN (AMMONIA)) (APHA, 2017). The second-derivative Ultra-
Cerisola., 2010) which have shown promising results for the pollutants violet Spectrophotometric method was used for nitrate analysis, using
removal. Based on the thorough literature survey, none of the previously UV-VIS Spectrophotometer (Method name:4500-NO-3 NITROGEN (NI-
reported works had addressed the comprehensive treatment of urban TRATE)) (APHA, 2017). UV-1800 by Shimadzu Scientific Instruments
wastewaters focusing on the simultaneous removal of COD, nutrients, Inc. was used for the above experiments. Active chlorine was measured
coliform bacteria, and micropollutants and, more specifically, with the according to the Iodometric method (APHA 4500-Cl B). The surface
inexpensive PbO2 anode. morphology and structure of the electrode layers were examined using a
In this context, the main objective of this work is to study the elec- scanning electron microscope (SEM, Quanta 400). Bruker D8 Discover
trochemical oxidation process in treating the two different urban AXS Powder X-ray Diffractometer was used to analyze the salt deposi-
wastewaters (primary effluent and secondary effluent) using Ti/b-SnO2/ tion on the cathode. The scan range (2θ) is between 0� and 90� . Ion
PbO2 anode. Ti metal is the base substrate and the purpose of Sb-SnO2 Chromatography (Thermo Scientific: Integrion HPIC) was used to
interlayer is to reduce the resistivity by promoting the passage of elec- quantify cations (column: CS 12A; 4 � 250 mm) and anions (column AS
trons from base Ti substrate to outer PbO2 layer and to increase the 18; 4 � 250 mm) in wastewaters. Estimation of �OH radicals, generated
electrode stability (Xia and Dai, 2018). The specific objectives include:
(i) To prepare and characterize the Ti//Sb-SnO2/PbO2 electrode for Table 1
conducting electrochemical oxidation experiments (ii) To assess the Characteristics of WW1 and WW2.
effectiveness of the prepared PbO2 electrode in terms of the reduction in Parameter WW1* WW2* Effluent discharge
COD, nutrients, coliform bacteria, and micropollutants during the standardsa
electrochemical oxidation of the two urban wastewaters and compare pH 7.05–7.6 7.35–8.1 6.5–9
the results (iii) To estimate the energy requirements for electrochemical Electrical conductivity 1.5–1.8 3.4–4.7 –
treatment of two wastewaters. (mS/cm)
COD (mg L 1) 240–456 60.8–152 < 50
Phosphorous (mg L 1) 5.5–10.51 12–13.9 5
2. Materials and methods Ammonia-nitrogen (mg 6.97–34.85 18.6–67.9 <5
L 1)
2.1. Chemicals and sample collection Nitrate Nitrogen (mg L 1) 1.58–3 0.62–1.5 10
Bacterial count (CFU/ml) 288800 221000 <100
Sodium sulfate (anhydrous), Potassium dichromate, Ammonium iron * WW1 - Wastewater sample collected from the influent of the SBR (Sequencing
(II) sulfate, Sulfuric acid, Hydrochloric acid were procured from Merck Batch Reactor) tank of the Sewage Treatment plant, IIT madras.
& Co., Inc. Individual reagent preparations was done according to pro- * WW2 - Wastewater sample collected from the artificial constructed wetland,
cedures described in APHA, 2017). Nutrient agar (HiMedia Laboratories Chennai Mathematical Institute, SIPCOT.
a
Effluent discharge standards (CPCB, 1986, 2015).

2
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

on the prepared electrode, was quantified according to the methodology 2.4. Gas chromatography and mass spectrometry (GC-MS) analysis
developed by Tai et al., 2004 using HPLC. Detailed conditions of HPLC
for �OH radical analysis was reported in the supplementary information, A detailed description of solid-phase extraction and GC-MS analysis
Text S1. Standard plate count test and ATP test were conducted on the for the micro-pollutants in the wastewaters can be found in the sup-
samples before and after treatment. The plating of all samples was plementary information, Text S1.
carried out inside a horizontal laminar flow in order to maintain sterile
conditions. The ATP was measured using the HygienaAquaSnap™ Total 3. Results and discussions
and AquaSnap free swabs placed in the HygienaEnSURE TM Lumin-
ometer which gave the ATP count in terms of RLU (Relative light units). 3.1. Physicochemical characterization of the electrode material
The microbial contamination in the wastewater samples was estimated
through the difference between the total and free ATP measurements. The morphology of the electrode layers was studied using SEM mi-
The heavy metal, lead was analyzed using Inductively Coupled Plasma crographs. As shown in Fig. 1a, the morphology of the pure titanium
Mass Spectrometer (ICP-MS) (PerkinElmer; Model: NexION 300X). substrate is even but with scratches observed on the surface. Fig. 1b
shows the roughened Ti surface after sandblasting. The rough structure
2.3. Experimental setup was well-formed and would provide a good bond with the interlayer.
The surface morphology of the antimony doped tin dioxide electrode is
2.3.1. Fabrication of Ti/Sb-SnO2/PbO2 electrode represented in Fig. 1c. The role of the inner Sb-SnO2 layer was to pro-
Titanium plates (5 cm � 2 cm x 0.1 cm and 5 cm � 5 cm x 0.1 cm) mote the better charge transfer from the innermost Ti layer to the outer
were used as electrode substrates and initially sandblasted. It was then PbO2 layer. This would reduce the electrode resistance and improve the
immersed in deionized water and ethanol and sequentially sonicated for lifetime of PbO2 layer. Fig. 1d is after electrochemical deposition of
15 min. The plates were then dipped in boiling HCl for 15 min (to PbO2. There are small crystals on the surface, which could efficiently
dissolve the oxide layer) and then placed immediately in deionized prevent the formation of TiO2 insulator layer with the Ti base. Crystal-
water so that the electrodes were not in contact with air. The interlayer line irregularities on the surface may be due to the difference in the
was prepared by thermal decomposition of Sb-SnO2 on the Titanium growth rate (Saratale et al., 2015). The compact crystal structure
plate. The Ti plate was immersed in a solution containing 20 g SnCl4. enhancement the conductivity and catalytic activity due to the improved
5H2O, 1 g SbCl3 in 100 mL propan-2-ol solution. After dipping at distribution of microcurrents (Weng et al., 2013). It was reported that
ambient temperature, the Ti plate was annealed in an oven at 100 � C for PbO2 is a water-insoluble form of lead (NCBI, 2019). The final electrode
0.5 h and in a muffle oven at 500 � C for 20 min. The above process was layer also provides a much larger specific surface area for increased
repeated eight times to fabricate the Ti/SnO2-Sb electrodes (Ma et al., electrochemical activity. EDX analysis of the electrode showed that the
2011). The outer layer of PbO2 was prepared by electrochemical anod- composition (mass %) of Pb and O to be 83.22% and 15.66%, respec-
ization (Hao et al., 2015). The process was conducted at 30 mA/cm2 for tively. The composition of fluorine (added as dopant) was 0.41% (Fig. S1
1 h in a water bath at the temperature 60 � 2 � C and powered by a DC and Table S2). The elements Ca, Mg and Cl, detected in trace amounts,
power supply (Scientific PSD 3003, 0–3A, 30V). The anode was could be the minor impurities from the external sources. The XRD
Ti/Sb-SnO2, and the cathode used was stainless steel. The electrolyte spectrum for the PbO2 electrode was presented along with the miller
was composed of 0.5 M Pb(NO3)2, 0.04 M NaF, and 0.1M HNO3 in 500 indices (Fig. S2). The diffraction peaks of the prepared anode are mostly
mL distilled water. During the anodization process, Pb2þ ions from lead in a match with a β-PbO2 crystalline structure and consistent with the
nitrate undergo oxidation to Pb4þ state and form crystals of PbO2 on the literature results (Cao et al., 2009; Dai et al., 2016).
electrode surface, as shown in Eqs. (4)–(7) below (Velichenko and
Devilliers, 2007). 3.2. Electrochemical characterization of the electrode material
At anode surface
Cyclic voltammetry (CV) experiments were performed with a 3-elec-
H2O → �OH þ Hþ þ e (4)
trode system consisting of Ti/Sb-SnO2/PbO2 as a working electrode,
Pb2þ þ �OH → Pb(OH)2þ (5) stainless steel as a counter electrode and Ag/AgCl as a reference elec-
trode. Fig. 2 represents the cyclic voltammetry for PbO2 electrode in
Pb(OH)2þ þ H2O → Pb(OH)2þ þ
2 þ H þ e (6) 0.1M Na2SO4 solution with a pH of 7.1 � 0.15. The overpotential for
oxygen evolution was found to be nearly 1.7 V V. Ag/AgCl electrode in
Pb(OH)2þ
2 → PbO2 þ 2H
þ
(7)
0.1 M Na2SO4 solution (without any contaminant). The value of the
Before using it for the experiments, the prepared Ti/Sb-SnO2/PbO2 oxygen evolution overpotential for the Ti/Sb-SnO2/PbO2 electrode is
electrode was thoroughly washed in distilled water. similar to the values (1.6 V) reported in the literature with the same
background electrolyte solution of Na2SO4 (Cao et al., 2009; Dai et al.,
2.3.2. Electrochemical treatment 2016). However, few studies reported an overpotential for oxygen
The electrochemical treatment was performed in a single cell reactor. evolution on PbO2 to be 1.9 V vs SHE, higher than that obtained in this
A glass beaker (capacity 500 mL) was used for this purpose with a Ti/Sb- study (Chaplin, 2014; Panizza and Cerisola. 2009). This could be mainly
SnO2/PbO2 anode and a stainless steel cathode. 250 mL of wastewater due to the differences in background electrolytes and reference elec-
sample was treated at a time, with stirring being provided by a 4 cm trodes used for CV tests. The reported value of 1.9 V V. SHE was obtained
magnetic stirrer (90–130 RPM). DC power supply was provided by in a highly acidic solution of 1 M H2SO4 whereas, in this present study,
Scientific PSD3003 (30V, 3A Single Power Supply). The experiments CV test was conducted in neutral salt solution of sodium sulfate. At
were carried out by varying parameters such as current density (10 mA/ anode surface, water undergoes oxidation reactions by Eqs. (1) and (3)
cm2, 30 mA/cm2, 50 mA/cm2), exposed electrode surface area (12 cm2 which are competitive. Owing to higher oxygen evolution overpotential
and 30 cm2), and reaction run time to find the optimum electrochemical as in the case of PbO2 electrode, electrochemical production of hydroxyl
setup. A detailed experimental plan can be found in the supplementary radicals would be facilitated by reaction Eq. (1). Higher the over-
information, Table S1. potential for oxygen evolution, lower will be the side reaction of oxygen
formation and the higher the oxidation efficiency of organic pollutants.
Unlike in conventional platinum and mixed metal oxide electrode sys-
tems, oxidation efficiencies of pollutants are relatively higher in PbO2
electrode systems due to the higher rates of reaction Eq. (1) than Eq. (3)

3
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Fig. 1. a) Titanium plate b) Titanium plate after sandblasting c) Antimony doped tin dioxide electrode layer d) Lead dioxide electrode layer.

3.3. Electrochemical treatment of wastewater

3.3.1. Raw water characterization


Various characteristics such as pH, electrical conductivity, chemical
oxygen demand (COD), phosphorous, ammonia-nitrogen, and nitrates
were analyzed for both the untreated wastewater samples. Table 1
shows the various physicochemical characteristics of untreated WW1
and WW2 samples. The physicochemical characteristics of the WW1
sample had pH in the range of 7.05–7.6, 1.5–1.8 mS/cm electrical
conductivity, 204–456 mg L 1 of COD, 5.5–10.51 mg L 1 of phospho-
rous, 6.97–34.85 mg L 1 of ammonia-nitrogen and 1.58–3 mg L 1 of
Nitrate nitrogen. Similarly, the characteristics of the WW2 sample
having a pH in the range of 7.35–8.1 were 3.4–4.7 mS/cm electrical
conductivity, 60.8–152 mg L 1 of COD, 12–13.9 mg L 1 of phosphorous,
18.6–67.9 mg L 1 of ammonia-nitrogen and 0.62–1.5 mg L 1 of nitrate
nitrogen. The effluent discharge standards for sewage treatment plants
in India read 6.5–9 for pH, 50 mg L 1 for COD, 5 mg L 1 for ammonia-
nitrogen, 10–20 mg L 1 for nitrate-nitrogen, 5 mg L 1 for phosphorous
Fig. 2. Cyclic Voltammetry curve for Ti/Sb-SnO2/PbO2 in 0.1 M Na2SO4. pH of (CPCB, 1986, 2015). By this treatment, these parameters have been
the solution: 7.1 � 0.15. brought to be within the discharge standards after which water can be
safely reused. The conductivity of WW1 was found to be low (1.8
(Chaplin, 2014). There was no visible formation of oxidation peaks mS/cm) to run the electrochemical experiments. Hence 25 mM of so-
during forwarding sweep (0–3.5 V). However, during reverse sweep dium sulfate was added in WW1 to increase the conductivity. WW2 had
(3.5–0 V), a reduction peak was observed between 0.75 and 0.9 V sufficient conductivity (3.4–4.7 mS/cm), and hence no salt addition was
(Fig. 2). This may correspond to the reduction of PbO2 to Pb3O4 form required. Ionic species such as Cl , SO24 , Ca2þ, Mg2þ, and Naþ were
(Cao et al., 2009). Since the curve follows almost the same path while it quantified, and their mean concentrations presented in Table 2. Both
returns to the initial potential (0 V–3 V to 0 V V. Ag/AgCl), it may also WW1 and WW2 contain relatively high chloride ion content (402.65 and
indicate that the working electrode is stable, even and does not possess 320 mg L 1 respectively) amongst all the ions. During electrochemical
any contaminants on its surface. oxidation, the chloride ions produce active chlorine species, which can
aid in the degradation of organics and disinfection.

4
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Table 2 min. The faster degradation of COD in WW2 could be mainly due to its
Ionic composition. low initial COD value (55–60 mg L 1), while the high initial COD value
Type of ion WW1 WW2 (~360 mg L 1) in WW1 resulted in slower degradation kinetics. It was
reported that (Silva et al., 2018; Sir� es et al., 2014) at high current
Cl (mg/L) 402.65 320
SO24 (mg/L) 43.82 94.03 applied currents there might be a decrease in organics removal rates due
Ca2þ (mg/L) 86.56 146.37 to the increased rates of possible parasitic reactions (O2 and H2 evolu-
Mg2þ (mg/L) 53.13 88.62 tions). However, in this study, at high currents, COD degradation ki-
Naþ(mg/L) 183.1 68.68 netics and removal percentages were positively improved (Fig. 4b and
d). The influence of the parasitic reactions specifically on COD removal
3.3.2. Hydroxyl radical estimation might not be very dominant at the current densities used in this study.
Hydroxyl radical estimation was carried out to determine the oxidant The nature of COD degradation kinetics was observed to be different
availability for organics degradation at the electrode of interest. Fig. 3 for WW1 and WW2, which was primarily attributed to the differences in
represents the concentration of hydroxyl radicals (OH� radicals) pro- wastewater quality (Table 1) that can be explained as follows. WW1 was
duced (in-situ) on the Ti/Sb-SnO2/PbO2 anode. Since the lifetime of the a high strength primary effluent (COD ~360 mg L 1), and the nature of
OH� radicals are very less, the addition of DMSO (CH3)2SO) acts as a COD was a highly biodegradable type. WW2 was a secondary effluent
trapping agent for �OH radical and produce formaldehyde (HCHO) ac- with low strength (COD: 55–60 mg L 1), and the nature of COD could be
cording to Eqs. (8)–(10) (Babbs and Griffin, 1989). a more non-biodegradable type. The COD degradation kinetics can also
be affected significantly by the presence of chloride ions in wastewater.
� OH þ (CH3)2SO → CH3SO2H þ �CH3 (8) It was well reported that chloride ions undergo oxidation at the anode
and yields active chlorine species such as Cl2 HOCl and OCl (Moreira
(9)
et al., 2017; Mostafa et al., 2018; Panizza and Cerisola, 2009) according

CH3 þ O2 → CH3OO �

2CH3OO� → HCHO þ CH3OH þ O2 (10) to Eq. (11) to Eq. (13)

HCHO is a stable product and doesn’t undergo degradation with �OH 2Cl- → Cl2 þ 2e (11)
radicals. Measurement of HCHO concentration would indirectly repre-
Cl2(aq) þ H2O → HClO þ Cl þ H þ
(12)
sent the total amount of �OH radicals produced in the reactor. It was
observed that the concentration of �OH radicals increased with time as a HClO ↔ ClO þ H þ
(13)
close-linear trend. It indicates that the rate of formation of �OH radical
was almost the same for the whole treatment duration of 2 h. At the end These active chlorine species can accelerate the removal of organic
of 2 h, the concentration of �OH radicals were observed to be 9.35E-05 species and abatement of COD (Radjenovic and Petrovic, 2017). The
M. differences in chloride concentrations in WW1 and WW2 might also
contribute to different COD removal kinetics observed in both waste-
3.3.3. COD removal waters. Electrogenerated active chlorine (Cl2) concentrations were
The COD degradation studies were carried out in both the waste- measured in WW1 and WW2 for different current densities, as shown in
water samples. For WW1 and WW2, different current densities of values Table 3. At 30 mA/cm2, the active chlorine levels were found to be
10 mA/cm2 and 30 mA/cm2 were applied to investigate the degradation relatively low in WW1 (5.3 mg L 1) than in WW2 (22.9 mg L 1) after the
of COD for an exposed electrode area of 12 cm2. Fig. 4b shows the complete removal of the COD. Since WW1 had high COD, the amount of
different values of COD at 0 h, 2.5 h, and 5 h for the different current Cl2 consumed for the oxidation of organics would be more, resulting in
densities. It can be seen from the graph that, at 10 mA/cm2 current low residual levels of active chlorine. On the other hand, WW2 was a low
density, the COD removal was found out to be 21% and 35% at hydraulic COD type, which requires less amount of Cl2 for the oxidation of or-
retention time (HRT) of 2.5 h and 5 h, respectively. Similarly, at 30 mA/ ganics and hence detected with relatively high residual active chlorine
cm2, the removal was found out to be 43% and 58% at HRT of 2.5 h and levels. It was also observed that the concentration of chloride ions (Cl )
5 h respectively. In the case of WW2, it can be seen from the graph was decreased (For e.g., 402.65 to 355.55 mg L 1 at 30 mA/cm2 in
(Fig. 4d) that, at 10 mA/cm2, the COD removal was 42% at 30 min, and WW1) during the electrochemical treatment for all the conditions
there was ~99.9% COD removal at 1 h. At a current density of 30 mA/ (Table 3) which may further support the consumption of Cl ions to form
cm2, it was observed that ~99.9% of COD was removed in less than 30 active chlorine species. From the above results (Fig. 4), it can be deduced
that the COD removal percentage increases with the increase in the
applied current density, and it also increases with an increase in the
hydraulic retention time. This may be explained by, as the current
density increases, the production of hydroxyl radicals and active chlo-
rine species also increase which in turn increases the degradation rates
of the pollutants present in water (Li et al., 2017; Panizza and Cerisola,
2004; Sanchez et al., 2018; Ganiyu et al., 2017).
For both WW1 and WW2, the COD degradation was also studied for
two different electrode areas of 12 cm2 and 30 cm2 at an applied current
density of 30 mA/cm2. In the case of WW1, for an electrode area of 12
cm2, the COD removal was only 58% after 5h, and for an electrode area
of 30 cm2, the removal was increased to 99% for the same HRT (Fig. 4a).
In the case of WW2, for an electrode area of 12 cm2, there was ~99.9%
COD removal at the end of 60 min, whereas, for an electrode area of 30
cm2, the same removal was attained in 30 min (Fig. 4c). An increase in
the surface area had improved the COD degradation kinetics by nearly 2
times for both the urban wastewaters, which means the treatment du-
rations can be shortened significantly by 50% for the same COD levels.
Fig. 3. Concentration of hydroxyl radicals generated on Ti/Sb-SnO2/PbO2 The results could be associated with the increase in the production of
anode. Conditions: 0.1M Na2SO4, reactor volume: 250 ml; anode and cathode hydroxyl radicals and active chlorine species because the active sites for
surface area: 12 cm2, Applied current density: 30 mA/cm2.

5
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Fig. 4. (a) and (c) Effect of electrode area on COD degradation kinetics under operating conditions of 30 mA/cm2 current density and detention time of 5 h and 1 h
for WW1 and WW2 respectively. (b) and (d) Effect of current density on COD degradation kinetics under operating conditions of 12 cm2 and 30 cm2 and detention
times of 5 h and 1 h for WW1 and WW2 respectively.

reasonable fit to predict COD degradation by electrochemical oxidation


Table 3
on PbO2 anodes.
Concentrations of chloride ion and active chlorine before and after the
treatment.
3.3.4. Phosphorous removal
Waste Current Condition Chloride Active
The phosphorus removal was measured before and after the elec-
water density(mA/ (Cl ) (mg/L) chlorine (mg/
cm2) L)
trochemical treatment for both the wastewater samples. For WW1, it
was studied by varying three different parameters: current density- 10
WW1- 402.65 0
mA/cm2, 30 mA/cm2, 50 mA/cm2; hydraulic retention times (HRTs) -3h
– –
Raw
WW1- 30 25 mM Na2SO4, 355.55 5.3 and 5h; electrode area- 12 cm2 and 30 cm2. Fig. 5a shows the initial and
Treated 12 cm2, 5h final value of phosphorous for the different operating conditions for
WW1- 10 25 mM Na2SO4, 369.55 5.1 WW1. It can be seen from the graph that, using an electrode of area 12
Treated 12 cm2, 5h
cm2 at the current density of 30 mA/cm2, the phosphorus removal was
WW2 - 320.00 0
found to be nearly 36% and 43% at HRTs of 3 h and 5 h respectively.
– –
Raw
WW2- 30 12 cm2, 1 h (no 281.29 22.9 However, for an electrode area of 30 cm2 and for electrolysis time of 5 h,
Treated external salt the % removal was found out to be 88% and 50% for current densities
WW2- 10 12 cm2, 1 h (no 298.99 4.8 10 mA/cm2 and 30 mA/cm2 respectively. It can also be seen that for an
Treated external salt
HRT of 5h, using a 12 cm2 electrode, the removal was determined to be
72% and 87% at current densities 50 mA/cm2and 10 mA/cm2 respec-
the reaction increased when the anode area was increased (Huang et al., tively. In the case of WW2, removal was studied by varying four
2018). The rate of hydroxyl radical production in the system was different parameters: current density- 10 mA/cm2 and 30 mA/cm2;
assumed to be a constant, and the rate of change of COD may follow HRTs- 5h, 2.5h and 1 h; electrode area- 12 cm2 and 30 cm2 and elec-
pseudo-first-order kinetics. Consequently, Table S3 presents the esti- trolyte addition. It can be seen from the graph that, for a 30 cm2 elec-
mated first-order rate constants and r2 values for the COD degradation trode area at HRT of 1 h, the removal was found out to be 98% and 99%,
rates for both the wastewater samples for different operating conditions at current densities 10 mA/cm2 and 30 mA/cm2 respectively. Also, for a
of current density and electrode area. The range of r2 values (0.85–0.99) 12 cm2 electrode, at a current density of 30 mA/cm2, the removal was
indicated that a pseudo-first-order kinetic equation could be a determined as 99% and 97% for HRTs of 2.5 and 1 h, respectively.

6
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

precipitation rate (EPR) of phosphate. However, an increase in the


surface area had shown a positive effect on phosphorus removal in
WW1. On the other hand, In WW2, an increase in applied currents did
not reduce the phosphate removal percentage (Fig. 5b), as previously
observed in the case of WW1. This contrasting behaviour within two
wastewaters could be attributed to differences in wastewater qualities.
WW1 contains high COD with low conductivity and also low initial
phosphate concentration. WW2 was a low COD with high conductivity
and phosphates concentration. EPR of phosphates in WW1 could have
slowed down under the influence of high background COD and low
initial phosphates concentration; while EPR of phosphates in WW2 was
not influenced by the low background COD. Due to high initial con-
ductivity and phosphate concentrations, there was a better removal and
recovery of phosphate in WW2.
Phosphorous removal can be further supported based on the concept
of solubility product (Ksp) of the compound. The compound 4CaO.P2O5
is a composite precipitate containing two compounds CaO and P2O5, and
with appropriate stoichiometric proportions and conditions, it can form
Ca3 (PO4)2 (calcium phosphate) as per the below reaction Eq. (14)

3CaO (s) þ P2O5 (s) → Ca3(PO4)2 (14)


16
The Ksp values for calcium phosphate in WW1 (6.73 � 10 ) and
WW2 (8.51 � 10 15) were found to be greater than the equilibrium Ksp
value (2.07 � 10 33), indicating the supersaturated conditions with
potential for precipitation of phosphate salts. Although the wastewater
was supersaturated with [Ca2þ] and [PO34 ] ions, chemical precipitation
was not observed visibly. Only during the electrolysis time, electro-
chemical precipitation was dominant, and there was a deposition of
phosphate salt on the cathode surface. This is because the electro-
chemical process causes electron mobility, which induces the formation
of the precipitate. Electrochemical recovery of phosphate from real
wastewaters was demonstrated in a few recent studies. Cid et al. (2018)
reported the recovery of phosphate as hydroxyapatite on the stainless
steel cathode from human wastewater. In the study by Lei et al. (2017),
the P concentration was shown to be reduced from 8 to 2.3 mg l 1 on
titanium cathode by the formation of calcium and magnesium phos-
phates, as observed in our study. These salts were appeared to be coated
on the cathode surface as a thick precipitate (Fig. S3). The voltage versus
time graph (Fig. S4) indicated that there was no major variation in
output cell voltage for constant current input. This indicates that the
precipitation on the cathode surface did not bring the major change in
electrochemical cell resistance during 5 h of operation as these salts are
highly conductive in nature. After every cycle, the phosphate salts were
recovered from the cathode surface by mechanical scraping. The final
values of P in treated waters were observed to be satisfying the discharge
Fig. 5. (a) Phosphorous removal under different experimental conditions of 30 limit of 5 mg L 1 of phosphorus (Fig. 5).
mA/cm2, 10 mA/cm2 and 50 mA/cm2 current densities, detention times of 3 h
and 5 h and electrode areas of 12 cm2 and 30 cm2 for WW1. (b) Phosphorous
3.3.4.1. X-ray diffraction (XRD) and SEM-EDS analysis of cathodic pre-
removal under different experimental conditions of 30 mA/cm2 and 10 mA/cm2
current densities, detention times of 5 h, 2.5 h and 1 h and electrode areas of 12 cipitate. Fig. 6 represents the XRD patterns of the solid precipitate
cm2 and 30 cm2 for WW2. deposited on the stainless steel cathode after the electrochemical treat-
ment of WW1 and WW2. In the case of WW1, the XRD spectrum shows 4
major diffraction intensity peaks of 100%, 63%, 26% and 34% at 2θ
Similarly, for a 12 cm2 electrode, at 10 mA/cm2, the removal was found
values of 29.29 ο, 37.82 ο, 43.05 ο, and 47.39 ο respectively. Each peak
out to be 99% at the end of 5 h. At 30 mA/cm2 current density, using an
was assigned to different compounds with the highest peak being
electrode area of 12 cm2, for an HRT of 2.5 h, under with and without
assigned to calcium oxide (CaO). The second highest peak was assigned
salt conditions, the removal % was found to be 99% and 98% respec-
to tetra calcium phosphate (4CaO.P2O5) with an intensity of 63%. The
tively. Similarly, for the same conditions as above with the HRT being 1
other compound which was assigned having an intensity of 34% was
h, the percentage removal was determined to be 97% and 98% with and
calcium dioxide (CaO2). In the case of WW2, the X-Ray diffraction
without salt conditions.
spectrum shows 5 major diffraction intensity peaks of 100%, 14%, 29%,
The reason for this phosphorous removal can be explained by the
26% and 14% at 2θ values of 29.29 ο, 35.78 ο, 39.31 ο, 43.05 ο, and 47.40
electrochemically induced precipitate (calcium and magnesium salts)
respectively. In this case, also, the highest peak was assigned to cal-
that was recovered from the cathode after the treatment process. In
ο

cium oxide. Tetracalcium phosphate, magnesium pyrophosphate, and


WW1, at low current densities, higher removal of phosphorus was
calcium dioxide were assigned to the peaks at 2θ values of 35.781 ο,
observed than at high currents (Fig. 5a). This may be explained as fol-
43.002 ο, and 47.334 ο respectively. Miller indices for the major
lows: At high currents, the high rates of competitive reactions (e.g. H2
diffraction peaks of two samples were also provided in Table S4.
gas evolution) at cathode might have affected the electrochemical

7
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

setups, the increase in nitrate concentrations was found to be around 7


ppm for the smaller electrodes. For the 50 mA/cm2 setup, the increase
was found to 7.5 ppm. The maximum increase was observed for the 10
mA/cm2, 5 h, 12 cm2 and the 30 mA/cm2, 5 h, 30 cm2setup, the increase
being 16.4 and 16.3 ppm respectively. However, the 10 mA/cm2,5 h, 30
cm2 setup only had an increase of 1.52 ppm. In the case of WW2, for the
30 mA/cm2, 12 cm2, 2.5 h setups, the increase in nitrate concentration
was 2.77 ppm when 25 mM Na2SO4 salt was added and 3.05 ppm when
salt was not added (Fig. 7d). For the 30 mA/cm2, 12 cm2, 1 h setup in
Fig. 7d, the increase in nitrate concentration was 1.8 ppm when salt was
added and 0.76 ppm when salt was not added. At the experimental
conditions of 10 mA/cm2, WW2, 5 h, 12 cm2, with salt had an increase of
4.48 ppm. Similarly, in the case of WW2, 10 mA/cm2, 1 h, 30 cm2,
without salt also showed an increase of 2.2 ppm. The maximum increase
amongst the 30 mA/cm2 setups was seen for the larger electrode at 1 h,
which was 8.76 ppm. For WW1, nitrate formation had not shown a
definite trend with applied current density or surface area of the elec-
trode. But for WW2, the end nitrate levels were found to be low at 10 and
30 mA/cm2 applied current densities with 12 cm2 surface area electrode.
Zorpas (2011) observed that for pH below 8, NH3-N was oxidized to
nitrates and for pH above 8 it is transformed mainly to nitrogen gas. The
pH of both WW1 and WW2 was observed to be between 7 and 8 in all
cases before treatment. It only slightly increased after treatment.
It was noted that for most of the conditions, increase in nitrate
Fig. 6. XRD patterns of the salt deposited at stainless steel cathode after the
electrochemical treatment of WW1 for 5 h at 10 mA/cm2 and WW2 for 1 h at concentration fell below the permissible discharge limits (10 mg L 1 of
10 mA/cm2. Nitrate-N) for WW2 (Fig. 7d) whereas the final nitrate levels were higher
than or just equal to permissible limits for WW1 (Fig. 7b). Also, initial
concentrations of NH3 were relatively higher in WW2. The above ob-
SEM-EDS mapping was carried out to determine the elemental
servations indicated that NH3 conversion to nitrate might have exhibited
composition of the cathodic precipitate (Table 4). In cases of both
dependence on the quality of wastewater being treated. The electro-
wastewater samples, the major salts deposited were found to be calcium
chemical method of NH3 removal may occur by (Garcia-Segura et al.,
and magnesium oxides owing to their highest mass % in the precipitate.
2018b, 2019) (i) conversion to NO3 by direct oxidation at anode, (ii)
The mass% of oxygen, calcium, and magnesium was found out to be
scavenging of NH3 by chlorine species that can produce N2 and, (iii)
37.90%, 7.92%, 15.88% in WW1 and 42.08%, 18.65% and 8.82% in
formation of chloramines by the reaction between NH3 in wastewater
WW2 respectively. An example EDX spectrum for WW1 was presented in
and electrogenerated active chlorine species (Cl2, HOCl) at anode ac-
Fig. S5. The mass % of phosphorus was found to be 2.19% and 0.32% in
cording to Equations (15)–(17) below (Crittenden et al., 2012)
WW1 and WW2, respectively. Traces of chlorine (0.61% and 0.01%
were also determined in the precipitate deposited in the cathode. NH3 þ HOCl → NH2Cl þ H2O (15)

3.3.5. Removal of ammonia-nitrogen and nitrates NH2Cl þ HOCl → NHCl2 þ H2O (16)
The presence of ammonia-nitrogen (NH3-N) causes odour problems
NHCl2 þ HOCl → NCl3 þH2O (17)
and affects disinfection efficiency. During chlorination, NH3 concen-
trations greater than 0.2 mg L 1 can consume up to 68% of the chlorine Some of the nitrogen loss can also occur by the production of NOx
(WHO, 2003). The initial and final ammonia-nitrogen and species during electrochemical oxidation as found by Garcia-Segura
nitrate-nitrogen concentrations were measured before and after elec- et al. (2017). The NH3, in the presence of active chlorine species, can
trochemical treatment. For WW1, the NH3-N concentration was between undergo oxidation to yield N2 gas or NO3 according to Equations (18)
6.97 and 34.85 mg L 1 and for WW2, it was between 26.3 and 67.9 mg and (19) given below.
L 1. In all cases, the removal percentages of ammonia-nitrogen were
found to be 100%, as seen in Fig. 7a and c respectively. 3HOCl þ 2NH3 → N2 þ 3H2O þ 3HCl (18)
There was an increase in nitrate concentrations for all the cases, as þ
4HOCl þ NH 3 → H þ NO3 þ H2O þ4HCl (19)
nitrate-N could be produced by the oxidation of ammonium-N. It can
also be reduced into nitrogen gas or ammonium-N on the cathode sur- The process of electro chlorination at the anode (that can produce
face (Deng et al., 2019). In Fig. 7b, all the cases involve the addition of active chlorine species) can greatly enhance the selectivity of NH3
salt as the conductivity of WW1 was very low. For the 30 mA/cm2 oxidation towards N2 gas evolution (Eq. (18)) up to ~100%. The re-
sidual nitrate, during the electrochemical treatment, can reduce to N2
gas (Eq. (20)) or back to NH3 (Eq. (21)) (Garcia-Segura et al., 2018b).
Table 4
However, in this study, NH3 was found to be completely removed
Elemental composition of the precipitate deposited after electrochemical
indicating that Eq. (21) was highly unselective during the electro-
treatment of WW1 and WW2 detected by SEM-EDS analysis.
chemical oxidation of domestic wastewaters.
Element Mass (%)

WW1 WW2
2 NO3 þ 12Hþ þ 10e → N2 þ 6 H2 Eo ¼ 1.17 V vs SHE (20)

Oxygen 37.90 42.08 þ


NO3 þ 9H þ 8e → NH3 þ 3H2 E ¼ �
0.12 V vs SHE (21)
Magnesium 15.88 8.82
Calcium 7.92 18.65 The electrochemical reduction of produced nitrate (Fig. 7) may be
Phosphorous 2.19 0.72 selectively enhanced by the use of alternative cathode materials in place
Sodium 1.49 0.32
of stainless steel used in this study. Some of the cathodes that had shown
Chlorine 0.61 0.01

8
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Fig. 7. (a) Degradation of NH3-N for WW1 under different operating conditions (b) Change in NO3-N for WW1 under different operating conditions (c) Degradation
of NH3-N for WW2 under different operating conditions (d) Change in NO3-N for WW2 under different operating conditions.

the better selectivity of nitrate towards N2 gas (Eq. (20)) were reviewed reduction in the bacterial count is mainly because of the inactivation of
to be copper alloys, tin, nickel, platinoid alloys and boron-doped dia- these bacterial cells by the electrogenerated reactive oxygen species.
mond (Garcia-Segura et al., 2018b). These cells may be subjected to two types of damages: the disinfectants
(hydroxyl radicals or the chlorine species) can react with the cell surface
3.3.6. Electrochemical disinfection components and can cause changes to the cell permeability; the second
The electrogenerated �OH radicals and active chlorine species (Cl2, type of damage may be due to the impairment of intracellular compo-
HOCl, and OCl ) not only helps in the removal of COD and other nu- nents such as the loss of DNA integrity (Huang et al., 2016; Ghernaout,
trients but also play an important role in disinfecting the coliform bac- 2017). The hypochlorite ions that are formed by the oxidation of chlo-
teria present in the wastewater by electrochemical oxidation. �OH rides are capable of eliminating the microorganisms present in the
radicals having high standard oxidation potential (2.8 V) can oxidize the wastewater (Dbira et al., 2019). Fig. 8b & d shows the reduction of ATP
bacterial cells leading to complete inactivation of them. Hence disin- count with respect to time in WW1 and WW2. The initial ATP count in
fection studies were carried out to determine the viable cells in both WW1 and WW2 are 3336 RLU and 6606 RLU. It can be seen from the
wastewater samples. This was attempted by measuring the Colony graph that ATP count becomes zero for HRTs 30 min and 6 min,
Forming Unit (CFU) and Adenosine Triphosphate count (ATP) in the respectively, for WW1 and WW2. The ATP count keeps decreasing by the
wastewater samples. Fig. 8a & c shows the degradation of CFU with subsequent increase in hydraulic retention time. The decrease in ATP
respect to time for both the wastewater samples. The CFU count for count could be due to changes in the growth rate of living cells. Thus,
untreated WW1 and WW2 was found to be 288800 CFU/mL and 221000 ATP count, a reliable measure of microbiological biomass, declines with
CFU/mL respectively. While it shows a decreasing behaviour with CFU count (Schneider and Gourse, 2004).
increasing HRT, it can be seen that the living cells were completely
inactivated, within 30 min of treatment time in the case of WW1 and 6 3.3.6.1. Study of bacterial cells morphology. SEM analysis was carried
min for WW2, the operating parameters being 30 mA/cm2 and area of out to understand the variation in the bacterial cell structure before and
the electrode of 12 cm2. Also, it can be seen that the log CFU/mL values after the application of electrochemical treatment. Fig. 9a represents the
have also gone down and become zero within HRTs of 30 min and 6 min morphology of the bacteria present in the raw wastewater sample. It
for wastewater samples 1 and 2 respectively. The main reason for this shows the presence of intact, rod-shaped and smooth surface bacterial

9
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Fig. 8. (a) & (c) Colony forming unit (CFU) and log CFU kinetics for WW1 and WW2 respectively (b) &(d) ATP count kinetics for WW1 and WW2 respectively.

cells with the average size of ~2 μm. Fig. 9b, c, and 9d represent the considered as the time taken for maximum COD removal (before satu-
alteration in the morphology of the bacterial cell structure after elec- ration) and not the total electrolysis time. The electrical energy con-
trochemical treatment times of 15 min, 30 min, and 1 h, respectively. At sumption (EC) for the treatment of the two wastewaters for different
15 min of treatment, slight deformities in the morphology of the bac- exposed electrode surface areas was calculated as shown in Table 5. For
terial cells were observed with the wrinkling of some bacterial cells, WW1, after 4 h of electrolysis, the energy consumption was 60.71 kWh/
which is mainly due to the breakage of the bacterial cell membrane by m3 for the smaller electrode surface area and 216 kWh/m3 for the larger
the attack of �OH radicals (Fig. 9b). In Fig. 9c, the formation of micro- electrode surface area. For WW2, after an hour of electrolysis, the energy
pores on cell membranes and the extensive wrinkling of cells were consumption was 21.88 kWh/m3 for the smaller electrode surface area
observed as the exposure of bacterial cells to �OH radicals tend to in- and 35 kWh/m3 for the larger electrode surface area after 30 min
crease with the increase of electrolysis time (Singh et al., 2016). Fig. 9d electrolysis time. Though the process is slightly energy-intensive, it can
shows the irregularities of the bacterial cells due to the enlarged mi- offer attractive solutions for specific challenges such as combined
cropores, which resulted in the shrinkage of cells due to the oozing out of destruction of refractory pollutants, combined secondary and tertiary
internal cell components at the end of electrolysis time of 1 h (Singh treatment, and complete disinfection. This is a process that addresses the
et al., 2016). This image shows the complete destruction and inactiva- removal of multiple parameters as we are applying both secondary
tion of bacterial cells owing to the deformities developed on the bacte- (COD, ammonia-nitrogen removal) and tertiary treatment (disinfection,
rial cell membrane and leakage of internal bacterial cell components due micropollutants, phosphorous removal) in a single process. Energy re-
to the oxidative cell damage. quirements of various other AO and biological processes were listed in
Table S5 for comparison.

3.4. Energy consumption


3.5. Analysis of micropollutants during treatment
Energy consumption indicates whether the implementation of elec-
trochemical technology is feasible or not. It gives us an idea of the GC-MS analysis can provide details of individual organic compounds
performance of the anode material and operating cost of the system. present in wastewater and their reduction in concentrations during the
Energy consumption per volume of treated effluent is expressed in kWh/ electrochemical treatment. Fig. S6 illustrates the overlaid GC-MS chro-
m3. It was calculated as per Eq. (22) (Gargouri et al., 2014). matograms of treated and untreated samples for WW1 and WW2. This
overlay can be used to compare the abundance of organic compounds in
Energy Consumption ¼
UIt
(22) wastewater before and after treatment. The y-axis represents the abun-
3600V dance and the x-axis denotes retention time. Each peak in the chro-
Where, U is the average cell voltage (V), I is the current (A), t is the matograms represents an organic compound in wastewater. It was
electrolysis time (s) and V is the volume (L). The electrolysis time was observed that the abundance of peaks for treated wastewater was

10
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

Fig. 9. SEM images showing morphology of bacterial cells on the application of electrochemical oxidation treatment at a current density of 30 mA/cm2 using
electrode area of 12 cm2 (a) 0 min sample (b) 15 min sample showing initial wrinkling of cells (c) 30 min sample showing extensive wrinkling of cells (d) 1 h sample
showing complete destruction of cells.

glycerol, bisphenol A and stearic acid. The complete details of com-


Table 5
pounds with their retention times, molecular weight, and chemical
Energy consumption.
formula were furnished in Tables S6 of Supplementary Information and
System Max. COD Current Average Energy S7. Table 6 presents the percentage-removal of the compounds in WW1
removal time Applied voltage (V) Consumption
and WW2 after electrochemical oxidation. It was observed that the
(h) (A) (kWh/m3)
emerging pollutants such as caffeine, bisphenol A and some phenolic
WW1, 30 4 0.36 10.54 60.71
compounds had undergone 100% removal whereas the acidic pollutants
mA/cm2,
5h, 12 cm2
such as benzoic acid, octanoic acid, decanoic acid exhibited greater than
WW1, 30 4 0.9 15 216 75% removal in WW1 and the percentage removals of glycolic acid,
mA/cm2,
5h, 30 cm2
WW2, 30 1 0.36 15.2 21.88
mA/cm2,
Table 6
1h, 12 cm2 Compounds in wastewater-1 and wastewater-2 and their percentage removals
WW2, 30 0.5 0.9 19.43 35 (Identified using GC-MS).
mA/cm2, S. Compounds in WW1 % S. Compounds in %
1h, 30 cm2 no Removal no WW2 Removal

1 Ethylene glycol 21.33% 1 3-Methylbutanoic 100%


significantly lesser than for untreated wastewater for WW1 and WW2 acid
2 Phenol 100% 2 Ethylene glycol 53.32%
samples (Fig. S6). This indicates that there was a considerable reduction
3 2-Propanol, 1-(2- 100% 3 Glycolic acid 100%
in the concentration of organic compounds by electrochemical oxida- methoxy-1-
tion. It can be seen from Fig. S6 that the abundance of compounds in methylethoxy)
WW2 was relatively lower than the abundance of WW1. This was due to 4 Hexanoic acid 70.7% 4 1- 100%
the fact that WW1 was received from primary effluent source and WW2 Cyclopentylethanol
5 Phenylethyl Alcohol 100% 5 Diethylene glycol 19.82%
from secondary effluent (after constructed wetlands during which there 6 Phenol, 2,3-dimethyl 100% 6 Octanoic acid 44.47%
would be some removal of organic compounds by biodegradation). 7 2-Isopropoxyphenol 100% 7 Glycerol 87.7%
Some of the pollutants, as identified by GC-MS, were ethylene glycol, 8 Benzoic Acid 82.87% 8 Bisphenol A 100%
phenol, phenylethyl alcohol, phenol 2,3-dimethyl-, benzoic acid, octa- 9 Octanoic acid 76.67% 9 Stearic acid 71.08%
10 Decanoic acid 79.20%
noic acid, diethyl phthalate, caffeine, decanoic acid in WW1. In WW2,
10 Diethyl Phthalate 87.85%
the pollutants were ethylene glycol, glycolic acid, diethylene glycol, 11 Caffeine 100%

11
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

octanoic acid, and stearic acid were 100%, 44.47%, and 71.08% ammonia-nitrogen and COD were observed to be 100% and 99% with
respectively in WW2 sample (Table 6). The relatively lower removal of respect to both samples, while phosphorous removal was 50% and 98%
acidic compounds was because they could be more stable and resistant for WW1 and WW2, respectively. An increase in current density
to oxidation as compared to pollutants such as caffeine, bisphenol A and increased the COD removal but did not increase the phosphorus
phenolic compounds. These organic compounds detected by GC-MS removal. As current density increases, �OH radical concentration in-
might have little correspondence to COD, as many compounds were creases that can cause higher degradation of COD. On the other hand, a
micropollutants. Previous studies have confirmed the reactivity of hy- decrease in phosphate removal at high currents was attributed to an
droxyl radicals with caffeine, bisphenol and acidic compounds such as increase in competitive reactions at the cathode (e.g., H2 gas evolution).
hexanoic acid, octanoic acid, and decanoic acid. Hydroxyl radicals can The phosphorus removal was majorly attributed to the formation of
effectively oxidize the caffeine in aqueous solutions and could produce calcium and magnesium phosphate precipitates on the cathode surface,
formic and oxalic acids as the byproducts. The mechanism of bisphenol which was confirmed by XRD and EDS analysis. This adds to the scope of
A degradation with �OH radicals involves mainly hydrogen abstraction, the potential to recover the phosphate as a resource from wastewater.
OH addition, and single electron transfer that could form the ketone and The electrochemical disinfection studies showed that complete disin-
polyhydroxylated stable compounds (Xiao et al., 2017). The rate con- fection was achieved within 30min and 6min for WW1 and WW2
stants of �OH radicals were reported to be increased with an increase in respectively. Complete destruction and morphological changes of bac-
the carbon number of naphthenic acids (NA) (e.g., hexanoic acid to terial cells were even confirmed by SEM analysis. A correlation between
decanoic acid). The reaction of �OH radical favours the abstraction of H ATP and bacterial colony count was investigated, which can be further
atom from C-H chain in NA compounds before the O-H bond breakage useful for disinfection kinetic studies. From the GC-MS analysis, com-
happens (Afzal et al., 2012). plete removal of certain micropollutants such as Caffeine, bisphenol A
and other phenolic compounds was observed. The energy consumption
3.6. Electrode stability for the entire process is reasonable as compared to other advanced
oxidation processes. The study in overall, addressed the complete
Fig. S7 (a and b) depicts the SEM micrographs of the Ti/Sb-SnO2/ removal of COD, NH3-N, coliform bacteria, few emerging pollutants and
PbO2 electrode before and after more than 60 h of usage. A slight change also demonstrated the better removal and recovery of phosphate.
in the surface morphology of the electrodes can be observed due to the However, certain issues still need to be addressed such as the presence of
formation of colloidal PbO2 particles that dissolve and redeposit on the chloride ions which may produce organo-chlorinated compounds in the
electrode surface (Pavlov and Monahov, 1996). However, the catalytic treatment process and the mechanism to periodically harvest phosphate
ability of the electrode was unaffected even after several hours of usage salts from the cathode surface which are also produced during the
and found suitable for a greater number of experimental runs. There was treatment.
no visible formation of inorganic salts deposition on the PbO2 anode
surface during all the experiments. Hence it was not required to pre-treat Declaration of competing interest
the PbO2 electrode after each experiment. Leaching of the lead (Pb) ions
from the PbO2 electrode was also evaluated after the electrochemical The authors declare that they have no known competing financial
oxidation of WW1 and WW2. The concentration of Pb was analyzed interests or personal relationships that could have appeared to influence
before and after the treatment using ICP-MS (Table S8) and operating the work reported in this paper.
conditions of ICP-MS were provided in Table S9. Initial concentrations
of Pb in WW1 and WW2 were estimated to be 8.20 μg/L and 14.07 μg/L, CRediT authorship contribution statement
respectively. After the completion of the electrochemical treatment, Pb
concentration was found to be 7.33 μg/L in WW1 and 8.16 μg/L in WW2 Bokam Rajasekhar: Conceptualization, Methodology, Investiga-
(Table S8). A slight reduction in the concentration of Pb can be ascribed tion, Writing - review & editing, Data curation, Project administration,
due to the electrodeposition/precipitation reaction occurred at the Funding acquisition. Urmika Venkateshwaran: Investigation, Writing
electrode surface. The measured values of Pb in both the treated waters - original draft, Validation. Nivetha Durairaj: Investigation, Writing -
was less than the Indian standard (10 μg/L) for drinking water original draft. Govindaraj Divyapriya: Conceptualization, Methodol-
(IS-10500, 2012; WHO, 2017). Analysis results show that there was no ogy, Project administration, Writing - review & editing, Funding
leaching of lead into treated waters from the electrode surface. More- acquisition. Indumathi M. Nambi: Conceptualization, Supervision,
over, only minimal variation in the pH of the treated samples was Project administration, Funding acquisition, Writing - review & editing.
observed, which also prevented the leaching of Pb. The near-neutral pH Angel Joseph: Writing - review & editing.
range of the wastewater samples did not affect the PbO2 electrode sur-
face. These results suggested that the prepared electrode (Ti/Sb-S- Acknowledgments
nO2/PbO2) is suitable and significantly stable to treat urban domestic
wastewaters. However, the stable behaviour of the Ti/Sb-SnO2/PbO2 Authors greatly acknowledge Industrial Consultancy and Sponsored
electrode may not be applicable for industrial wastewaters as the pH of Research, IIT Madras for funding the study through innovative project
the treated samples might become acidic or basic. scheme (2018–2019; Project no. CE3 2018), and also thankful to the
Department of Metallurgical and Materials Engineering, IIT Madras for
4. Conclusions providing the SEM facility. We would like to acknowledge Mr. Prince
Anantharaj, Technicican, Civil Engineering workshop, IIT Madras for his
In the present study, the applicability of an anodic oxidation process support in cutting and finishing of the Ti base electrodes. Authors are
in the treatment of two different urban wastewater samples (WW1 & thankful to Chennai Mathematical Institute, SIPCOT, Siruseri for
WW2) with Ti/Sb-SnO2/PbO2 as anode and stainless steel as a cathode providing wastewater samples.
was successfully evaluated. Different parameters were optimized sepa-
rately for both samples. Higher current density and larger electrode area Appendix A. Supplementary data
enhanced the degradation of the organic matter present in the waste-
waters. This could be mainly due to greater production of hydroxyl Supplementary data to this article can be found online at https://doi.
radicals at higher current densities and a large number of active sites org/10.1016/j.jenvman.2020.110469.
that promote enhanced oxidation of the organic pollutants at larger
electrode surface areas. Under the optimized conditions, removal of

12
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

References applications. Appl. Catal. B Environ. 236, 546–568. https://doi.org/10.1016/j.


apcatb.2018.05.041.
Garcia-Segura, S., Mostafa, E., Baltruschat, H., 2019. Electrogeneration of inorganic
Afzal, A., Drzewicz, P., Pérez-Estrada, L.A., Chen, Y., Martin, J.W., Gamal El-Din, M.,
chloramines on boron-doped diamond anodes during electrochemical oxidation of
2012. Effect of molecular structure on the relative reactivity of naphthenic acids in
ammonium chloride, urea and synthetic urine matrix. Water Res. 160, 107–117.
the UV/H2O2 advanced oxidation process. Environ. Sci. Technol. 46 (19),
https://doi.org/10.1016/j.watres.2019.05.046.
10727–10734. https://doi.org/10.1021/es302267a.
Gargouri, B., Gargouri, O.D., Gargouri, B., Trabelsi, S.K., Abdelhedi, R., Bouaziz, M.,
APHA (American Public Health Association), 2017. Standard Methods for the
2014. Application of electrochemical technology for removing petroleum
Examination of Water and Wastewater, 23rd edition. American public health
hydrocarbons from produced water using lead dioxide and boron-doped diamond
association, Washington, DC.
electrodes. Chemosphere 117, 309–315. https://doi.org/10.1016/j.
Babbs, C.F., Griffin, D.W., 1989. Scatchard analysis of methane sulfinic acid production
chemosphere.2014.07.067.
from dimethyl sulfoxide: a method to quantify hydroxyl radical formation in
Ghernaout, D., 2017. Microorganisms’ electrochemical disinfection phenomena. EC
physiologic systems. Free Radic. Biol. Med. 6, 493–503. https://doi.org/10.1016/
Microbiol 9, 160–169.
0891-5849(89)90042-7.
Hao, X., Wei, Y., Honghui, Y., 2015. Surface analysis of Ti/Sb-SnO2/PbO2 electrode after
Bian, X., Xia, Y., Zhan, T., Wang, L., Zhou, W., Dai, Q., Chen, J., 2019. Electrochemical
long time electrolysis. Rare Met. Mater. Eng. 44, 2637–2641. https://doi.org/
removal of amoxicillin using a Cu doped PbO2 electrode: electrode characterization,
10.1016/S1875-5372(16)60009-7.
operational parameters optimization and degradation mechanism. Chemosphere
Huang, X., Qu, Y., Cid, C.A., Finke, C., Hoffmann, M.R., Lim, K., Jiang, S.C., 2016.
233, 762–770. https://doi.org/10.1016/j.chemosphere.2019.05.226.
Electrochemical disinfection of toilet wastewater using wastewater electrolysis cell.
Cao, Jianglin, Zhao, Haiyan, Cao, Fahe, Zhang, Jianqing, ChunanCao, 2009.
Water Res. 92, 164–172. https://doi.org/10.1016/j.watres.2016.01.040.
Electrocatalytic degradation of 4-chlorophenol on F-doped PbO2 anodes.
Huang, K.L., Wei, K.C., Chen, M.H., Ma, C.Y., 2018. Removal of organic and ammonium
Electrochim. Acta 54 (9), 2595–2602. https://doi.org/10.1016/j.electacta.2008.10
nitrogen pollutants in swine wastewater using electrochemical advanced oxidation.
.049.
Int. J. Electrochem. Sci. 13, 11418–11431.
Chaplin, B.P., 2014. Critical review of electrochemical advanced oxidation processes for
IS-10500, 2012. Indian Standard Drinking Water–Specification (Second Revision).
water treatment applications. Environ. Sci.: Process. Impacts 16 (6), 1182–1203.
Bureau of Indian Standards (BIS), New Delhi.
https://doi.org/10.1039/c3em00679d.
Lei, Y., Song, B., van der Weijden, R.D., Saakes, M., Buisman, C.J.N., 2017.
Chiang, L.C., Chang, J.E., Wen, T.C., 1995. Indirect oxidation effect in electrochemical
Electrochemical induced calcium phosphate precipitation: importance of local pH.
oxidation treatment of landfill leachate. Water Res. 29, 671–678. https://doi.org/
Environ. Sci. Technol. 51, 11156–11164. https://doi.org/10.1021/acs.est.7b03909.
10.1016/0043-1354(94)00146-X.
Li, W., Yang, H., Liu, Q., 2017. Hydrothermal synthesis of PbO2/RGO nanocomposite for
Cid, C.A., Jasper, J.T., Hoffmann, M.R., 2018. Phosphate recovery from human waste via
electrocatalytic degradation of cationic red X-GRL. J. Nanomater., 1798706 https://
the formation of hydroxyapatite during electrochemical wastewater treatment. ACS
doi.org/10.1155/2017/1798706, 1 – 7.
Sustain. Chem. Eng. 6, 3135–3142. https://doi.org/10.1021/
Liu, Y.J., Hu, C.Y., Lo, S.L., 2019. Direct and indirect electrochemical oxidation of amine-
acssuschemeng.7b03155.
containing pharmaceuticals using graphite electrodes. J. Hazard Mater. 366,
CPCB (Central Pollution Control Board), 1986. General Standards for Discharge of
592–605. https://doi.org/10.1016/j.jhazmat.2018.12.037.
Environmental Pollutants Part-A : Effluents. The Environment (Protection) Rules,
Ma, C., Tan, F., Zhao, H., Chen, S., Quan, X., 2011. Sensitive amperometric
India. Available online at: http://cpcb.nic.in/GeneralStandards.pdf. (Accessed 12
determination of chemical oxygen demand using Ti/Sb–SnO2/PbO2 composite
June 2019).
electrode. Sensor. Actuator. B Chem. 155, 114–119. https://doi.org/10.1016/j.
CPCB (Central Pollution Control Board) ENVIS, 2015. News Letter on Control of
snb.2010.11.033.
Pollution. Issue-1 (Jan-Apr), India. Available online at: http://cpcbenvis.nic.in/envi
Ma, X., Wang, R., Guo, W., Yang, H., Liang, Z., Fan, C., 2012. Electrochemical removal of
s_newsletter/ENVIS%20Newsletter%20Jan%20-%20Apr%202015.pdf. (Accessed 12
ammonia in coking wastewater using Ti/SnO2þ Sb/PbO2 anode. Int. J.
June 2019).
Electrochem. Sci 7, 6012.
Crittenden, J.C., Trussell, R.R., Hand, D.W., Howe, K.J., Tchobanoglous, G., 2012.
Mandal, P., Gupta, A.K., Dubey, B.K., 2020. Role of inorganic anions on the performance
MWH’s Water Treatment: Principles and Design. John Wiley & Sons.
of landfill leachate treatment by electrochemical oxidation using graphite/PbO2
Dai, Q., Xia, Y., Sun, C., Weng, M., Chen, J., Wang, J., Chen, J., 2014. Electrochemical
electrode. J. Water Process Eng. 33, 101119. https://doi.org/10.1016/j.
degradation of levodopa with modified PbO2 electrode: parameter optimization and
jwpe.2019.101119.
degradation mechanism. Chem. Eng. J. 245, 359–366. https://doi.org/10.1016/j.
Martinez-Huitle, C.A., Ferro, S., 2006. Electrochemical oxidation of organic pollutants
cej.2013.08.036.
for the wastewater treatment: direct and indirect processes. Chem. Soc. Rev. 35,
Dai, Q., Zhou, J., Meng, X., Feng, D., Wu, C., Chen, J., 2016. Electrochemical oxidation of
1324–1340. https://doi.org/10.1039/b517632h.
cinnamic acid with Mo modified PbO2 electrode: electrode characterization, kinetics
Moreira, F.C., Boaventura, R.A., Brillas, E., Vilar, V.J., 2017. Electrochemical advanced
and degradation pathway. Chem. Eng. J. 289, 239–246.
oxidation processes: a review on their application to synthetic and real wastewaters.
Dbira, S., Bensalah, N., Ahmad, M.I., Bedoui, A., 2019. Electrochemical oxidation/
Appl. Catal. B Environ. 202, 217–261. https://doi.org/10.1016/j.
disinfection of urine wastewaters with different anode materials. Materials 12, 1254.
apcatb.2016.08.037.
https://doi.org/10.3390/ma12081254.
Mostafa, E., Reinsberg, P., Garcia-Segura, S., Baltruschat, H., 2018. Chlorine species
Deng, Y., Chen, N., Feng, C., Chen, F., Wang, H., Kuang, P., Feng, Z., Liu, T., Gao, Y.,
evolution during electrochlorination on boron-doped diamond anodes: in-situ
Hu, W., 2019. Treatment of organic wastewater containing nitrogen and chlorine by
electrogeneration of Cl2, Cl2O and ClO2. Electrochim. Acta 281, 831–840. https://
combinatorial electrochemical system: taking biologically treated landfill leachate
doi.org/10.1016/j.electacta.2018.05.099.
treatment as an example. Chem. Eng. J. 364, 349–360. https://doi.org/10.1016/j.
NCBI, 2019. Lead Dioxide: Compound Summary. National Centre for Biotechnology
cej.2019.01.176.
Information. Available online at: https://pubchem.ncbi.nlm.nih.gov/compound/Lea
Elaissaoui, I., Akrout, H., Grassini, S., Fulginiti, D., Bousselmi, L., 2019. Effect of coating
d-dioxide. (Accessed 11 September 2019).
method on the structure and properties of a novel PbO2 anode for electrochemical
Niu, J., Lin, H., Xu, J., Wu, H., Li, Y., 2012. Electrochemical mineralization of
oxidation of Amaranth dye. Chemosphere 217, 26–34. https://doi.org/10.1016/j.
perfluorocarboxylic acids (PFCAs) by Ce-doped modified porous nanocrystalline
chemosphere.2018.10.161.
PbO2 film electrodeEnviron. Sci. Technol. 46, 10191–10198. https://doi.org/
Fernandes, A., Santos, D., Pacheco, M.J., Ciríaco, L., Lopes, A., 2014. Nitrogen and
10.1021/es302148z.
organic load removal from sanitary landfill leachates by anodic oxidation at Ti/Pt/
Panizza, M., Cerisola, G., 2004. Electrochemical oxidation as a final treatment of
PbO2, Ti/Pt/SnO2-Sb2O4 and Si/BDD. Appl. Catal. B Environ. 148, 288–294. https://
synthetic tannery wastewater. Environ. Sci. Technol. 38, 5470–5475. https://doi.
doi.org/10.1016/j.apcatb.2013.10.060.
org/10.1021/es049730n.
Fernandes, A., Santos, D., Pacheco, M.J., Ciríaco, L., Lopes, A., 2016. Electrochemical
Panizza, M., Cerisola, G., 2009. Direct and mediated anodic oxidation of organic
oxidation of humic acid and sanitary landfill leachate: influence of anode material,
pollutants. Chem. Rev. 109, 6541–6569. https://doi.org/10.1021/cr9001319.
chloride concentration and current density. Sci. Total Environ. 541, 282–291.
Panizza, M., Cerisola, G., 2010. Applicability of electrochemical methods to carwash
https://doi.org/10.1016/j.scitotenv.2015.09.052.
wastewaters for reuse. Part 1: anodic oxidation with diamond and lead dioxide
Gaber, M., Abu Ghalwa, N., Khedr, A.M., Salem, M.F., 2013. Electrochemical
anodes. J. Electroanal. Chem. 638, 28–32. https://doi.org/10.1016/j.
degradation of reactive yellow 160 dye in real wastewater using C/PbO 2-, Pb.
jelechem.2009.10.025.
J. Chem. 2013, 9. https://doi.org/10.1155/2013/691763.
Pavlov, D., Monahov, B., 1996. Mechanism of the elementary electrochemical processes
Ganiyu, S.O., Oturan, N., Raffy, S., Esposito, G., Van Hullebusch, E.D., Cretin, M.,
taking place during oxygen evolution on the lead dioxide electrode. J. Electrochem.
Oturan, M.A., 2017. Use of sub-stoichiometric titanium oxide as a ceramic electrode
Soc. 143, 3616–3629.
in anodic oxidation and electro-Fenton degradation of the beta-blocker propranolol:
Radjenovic, J., Petrovic, M., 2017. Removal of sulfamethoxazole by electrochemically
degradation kinetics and mineralization pathway. Electrochim. Acta 242, 344–354.
activated sulfate: implications of chloride addition. J. Hazard Mater. 333, 242–249.
https://doi.org/10.1016/j.electacta.2017.05.047.
https://doi.org/10.1016/j.jhazmat.2017.03.040.
Garcia-Segura, S., Mostafa, E., Baltruschat, H., 2017. Could NOx be released during
Rao, A.N.S., Venkatarangaiah, V.T., 2014. Metal oxide-coated anodes in wastewater
mineralization of pollutants containing nitrogen by hydroxyl radical? Ascertaining
treatment. Environ. Sci. Pollut. Res. 21, 3197–3217. https://doi.org/10.1007/
the release of N-volatile species. Appl. Catal. B Environ. 207, 376–384. https://doi.
s11356-013-2313-6.
org/10.1016/j.apcatb.2017.02.046.
Rosal, R., Rodríguez, A., Perdig� on-Mel� on, J.A., Petre, A., García-Calvo, E., G�
omez, M.J.,
Garcia-Segura, S., Ocon, J.D., Chong, M.N., 2018a. Electrochemical oxidation
Agüera, A., Fern�andez-Alba, A.R., 2010. Occurrence of emerging pollutants in urban
remediation of real wastewater effluents—a review. Process Saf. Environ. 113,
wastewater and their removal through biological treatment followed by ozonation.
48–67. https://doi.org/10.1016/j.psep.2017.09.014.
Water Res. 44, 578–588. https://doi.org/10.1016/j.watres.2009.07.004.
Garcia-Segura, S., Lanzarini-Lopes, M., Hristovski, K., Westerhoff, P., 2018b.
Sanchez, A.S., Tejocote-P� erez, M., Fuentes-Rivas, R.M., Linares-Hern� andez, I., Martínez-
Electrocatalytic reduction of nitrate: fundamentals to full-scale water treatment
Miranda, V., Fonseca-Montes de Oca, R.M.G., 2018. Treatment of a textile effluent

13
B. Rajasekhar et al. Journal of Environmental Management 266 (2020) 110469

by electrochemical oxidation and coupled system electooxidation–salix babylonica. Velichenko, A.B., Devilliers, D., 2007. Electrodeposition of fluorine-doped lead dioxide.
Int. J. Photoenergy. https://doi.org/10.1155/2018/3147923. J. Fluor. Chem. 128, 269–276. https://doi.org/10.1016/j.jfluchem.2006.11.010.
Saratale, R.G., Hwang, K.J., Song, J.Y., Saratale, G.D., Kim, D.S., 2015. Electrochemical Weng, M., Zhou, Z., Zhang, Q., 2013. Electrochemical degradation of typical dyeing
oxidation of phenol for wastewater treatment using Ti/PbO2 electrode. J. Environ. wastewater in aqueous solution: performance and mechanism. Int. J. Electrochem.
Eng. 142, 04015064. Sci 8, 290–296.
Schneider, D.A., Gourse, R.L., 2004. Relationship between growth rate and ATP WHO, 2003. Ammonia in Drinking-Water: Background Document for Development of
concentration in Escherichia coli a bioassay for available cellular ATP. J. Biol. Chem. WHO Guidelines for Drinking-Water Quality. World Health Organization (No. WHO/
279, 8262–8268. SDE/WSH/03.04/01).
Silva, L.G., Moreira, F.C., Souza, A.A., Souza, S.M., Boaventura, R.A., Vilar, V.J., 2018. WHO, 2017. WHO Guidelines for Drinking Water Quality: First Addendum to the, fourth
Chemical and electrochemical advanced oxidation processes as a polishing step for ed. World Health Organization.
textile wastewater treatment: a study regarding the discharge into the environment Xia, Y., Dai, Q., 2018. Electrochemical degradation of antibiotic levofloxacin by PbO2
and the reuse in the textile industry. J. Clean. Prod. 198, 430–442. https://doi.org/ electrode: kinetics, energy demands and reaction pathways. Chemosphere 205,
10.1016/j.jclepro.2018.07.001. 215–222. https://doi.org/10.1016/j.chemosphere.2018.04.103.
Singh, R.K., Philip, L., Ramanujam, S., 2016. Disinfection of water by pulsed power Xia, Y., Wang, G., Guo, L., Dai, Q., Ma, X., 2020. Electrochemical oxidation of Acid
technique: a mechanistic perspective. RSC Adv. 6, 11980–11990. https://doi.org/ Orange 7 azo dye using a PbO2 electrode: parameter optimization, reaction
10.1039/C5RA26941E. mechanism and toxicity evaluation. Chemosphere 241, 125010. https://doi.org/
Sir�
es, I., Brillas, E., Oturan, M.A., Rodrigo, M.A., Panizza, M., 2014. Electrochemical 10.1016/j.chemosphere.2019.125010.
advanced oxidation processes: today and tomorrow. A review. Environ. Sci. Pollut. Xiao, R., Gao, L., Wei, Z., Spinney, R., Luo, S., Wang, D., Dionysiou, D.D., Tang, C.J.,
Res. 21, 8336–8367. https://doi.org/10.1007/s11356-014-2783-1. Yang, W., 2017. Mechanistic insight into degradation of endocrine disrupting
Tai, C., Peng, J.F., Liu, J.F., Jiang, G.B., Zou, H., 2004. Determination of hydroxyl chemical by hydroxyl radical: An experimental and theoretical approach. Environ.
radicals in advanced oxidation processes with dimethyl sulfoxide trapping and liquid Pollut. 231, 1446–1452. https://doi.org/10.1016/j.envpol.2017.09.006.
chromatography. Anal. Chim. Acta 527 (1), 73–80. https://doi.org/10.1016/j.aca. Zhu, X., Tong, M., Shi, S., Zhao, H., Ni, J., 2008. Essential explanation of the strong
2004.08.019. mineralization performance of boron-doped diamond electrodes. Environ. Sci.
Urtiaga, A., Fernandez-Castro, P., Gomez, P., Ortiz, I., 2014. Remediation of wastewaters Technol. 42, 4914–4920. https://doi.org/10.1021/es800298p.
containing tetrahydrofuran. Study of the electrochemical mineralization on BDD Zorpas, A.A., 2011. Alternative treatment of urban wastewater using electrochemical
electrodes. Chem. Eng. J. 239, 341–350. https://doi.org/10.1016/j.cej.2013.11.028. oxidation. Desalination Water Treat. 27, 268–276. https://doi.org/10.5004/
Vargas, R., Díaz, S., Viele, L., Nú~nez, O., Borr�
as, C., Mostany, J., Scharifker, B.R., 2014. dwt.2011.2028.
Electrochemical oxidation of dichlorvos on SnO2-Sb2O5 electrodes. Appl. Catal. B
Environ. 144, 107–111. https://doi.org/10.1016/j.apcatb.2013.06.016.

14

You might also like