1 s2.0 S0959652621023374 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Cleaner Production 314 (2021) 128119

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Effective removal of the heavy metal-organic complex Cu-EDTA from water


by catalytic persulfate oxidation: Performance and mechanisms
Qi Wang a, Yutong Li a, Yue Liu a, Jingyu Ren b, **, Ying Zhang c, Guangzhou Qu a, d,
Tiecheng Wang a, d, *
a
College of Natural Resources and Environment, Northwest A&F University, Yangling, Shaanxi Province, 712100, PR China
b
School of Petroleum Engineering and Environmental Engineering, Yan’an University, Yan’an, 716000, China
c
College of Information Science and Technology, Nanjing Forestry University, Nanjing, 210037, China
d
Key Laboratory of Plant Nutrition and the Agri-environment in Northwest China, Ministry of Agriculture, Yangling, Shaanxi, 712100, PR China

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

Handling editor: Prof. Jiri Jaromir Klemeš It is difficult to remove heavy metal-organic complexes from water by chemical precipitation because of the
strong complexation ability between heavy metal ions and organics. In this study, the removal of the Cu-
Keywords: ethylenediaminetetraacetic acid (Cu-EDTA) complex using autocatalytic persulfate (PS) oxidation was investi­
Persulfate gated. The Cu-EDTA removal efficiency reached up to 96.57% after 90 min of treatment by PS oxidation. A
Cu-EDTA
higher PS concentration favored Cu-EDTA removal; An increase in the initial concentration of Cu-EDTA benefited
Activation mechanisms
PS activation, and a greater removal performance was obtained at a lower Cu-EDTA initial concentration (0.1
Decomplexation
Removal mmol L− 1). Excessive Cu2+ accelerated Cu-EDTA removal, while superfluous EDTA suppressed it. Relatively
1
lower initial solution pH value favored Cu-EDTA removal. SO•-
4 , •OH, and O2 displayed significant roles in the
Cu-EDTA removal process, as they destroyed the chelating sites of the Cu(II) and EDTA molecules; finally small
molecular organic acids, alcohols, and NO−3 were produced. The released Cu(II) existed in the precipitates in the
forms of Cu-based carbonates, Cu-based hydroxides, and copper oxide. A possible decomposition pathway of Cu-
EDTA was proposed. Overall, multipathway activation of PS induced by heavy metal complexes could be an
effective technique for the removal of the heavy metal complexes.

1. Introduction Decomplexation is considered to be an essential step for effective


removal of the metal complexes (Zhao et al., 2015). In the decom­
Ethylenediaminetetraacetic acid (EDTA), an organic complexing plexation process, the heavy metal ions can be released, which provides
agent, is widely used in textiles (Lee et al., 2015), printing and dyeing convenience for subsequent recovery or precipitation of the free metal
(Kim and Baek, 2019), electroplating, and other industries because it can ions. Many oxidation methods, including Fe(III)/ultraviolet irradiation
well increase the activity of surfactants (Cuprys et al., 2018). As a result, (UV) (Shan et al., 2018), photoelectrocatalysis (Zhao et al., 2015),
heavy metal-EDTA complexes are easily detected in various water discharge plasma (Wang et al., 2019a; Liu et al., 2021), Fenton reaction
sources due to the powerful chelating ability of EDTA with heavy metals (Guan et al., 2018), and photocatalysis (Lan et al., 2018) have been used
(Lee et al., 2015). Compared with metal ions, the heavy metal complexes to remove the heavy metal complexes. Electro-Fenton combined with
are more toxic to environment and microorganisms (Ahamad et al., electrocoagulation was reported to be effective for Cu-EDTA elimination
2019). Arruda et al. (2010) reported that the half lethal concentration of from wastewater, whereas iron-bearing deposition in sludge needed
Cu-EDTA for Aedes aegypti was 32 mg L− 1. Removal of the heavy metal further treatment (Guan et al., 2018). Most nanophotocatalytic mate­
complexes has been a major challenge because they are extremely stable rials are composed of transition metals, which might lead to potential
over a wide pH range and strongly resistant to some conventional ecological risks to the environment; and the catalytic activity of the
removal methods (Wang et al., 2018). Thus, it has become a top priority photocatalysts is limited by effective light absorption and materials (Lan
to remove these complexes. et al., 2018). Non-thermal plasma oxidation could effectively remove

* Corresponding author. College of Natural Resources and Environment, Northwest A&F University, Yangling, Shaanxi Province, 712100, PR China.
** Corresponding author.
E-mail addresses: renjingyuyau@163.com (J. Ren), wangtiecheng2008@nwsuaf.edu.cn (T. Wang).

https://doi.org/10.1016/j.jclepro.2021.128119
Received 2 January 2021; Received in revised form 23 June 2021; Accepted 24 June 2021
Available online 25 June 2021
0959-6526/© 2021 Elsevier Ltd. All rights reserved.
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Cu-EDTA; however, electric power consumption limits its wide appli­ prepared by dissolving 1.2485 g CuSO4⋅5H2O into 100 mL deionized
cation (Wang et al., 2019b). Cu(II) in the Cu-EDTA could be replaced by water. EDTA stock solution with an initial concentration of 0.05 mol L− 1
Fe(III) to produce more stable Fe(III)-EDTA, and then the Fe(III)-EDTA was prepared by dissolving 1.8612 g EDTA-2Na into 100 mL deionized
was decomposed by subsequent UV to destroy the chelate (Shan et al., water. Cu-EDTA working solution was prepared by mixing the same
2018). volume of Cu2+ and EDTA stock solutions into 100 mL. The initial pH
Recently, persulfate (PS) oxidation, as an in-situ chemical oxidation value of 0.1 mmol L− 1 Cu-EDTA working solution was 4.0 without
technology, has received wide attention in pollutant removal (Pi et al., adjustment, which was comparable to the real electroplating wastewater
1
2019). SO•-4 , •OH, O2 , and O2 are probably generated by the activation
•-
(Fu et al., 2021). For the effects of initial Cu-EDTA concentration, 0.1,
of PS, which are responsible for organic pollutant degradation as re­ 0.2, 0.3, and 0.4 mmol L− 1 Cu-EDTA working solutions were prepared
ported by Dong et al. (2019), Pi et al. (2019), and Xu et al. (2020). The by mixing 200, 400, 600, and 800 μL Cu2+ and EDTA stock solutions into
activation of PS requires exogenous materials or an energy supply such 100 mL, respectively; the experiments were carried out at 10 mmol L− 1
as metal ions, heat, and UV irradiation. Cu-EDTA decomplexation and PS. For the effects of initial solution pH, 0.1 mmol L− 1 Cu-EDTA working
removal from water was achieved by UV/PS oxidation (Xu et al., 2017). solution was prepared by mixing 200 μL Cu2+ and EDTA stock solutions
The Co2+/PS oxidation system was proven to rapidly remove Cu-EDTA into 100 mL; then the initial solution pH was adjusted to 4, 6, 8, and 10
from wastewater (Xu et al., 2020). It was reported that heat combined using 0.5 mol L− 1 H2SO4 and NaOH, respectively; the experiments were
with Fe(II)-complexes could activate PS to remove hydrolyzed poly­ carried out at 10 mmol L− 1 PS.
acrylamide (Huang et al., 2020). It is expected that Cu-EDTA can be used For the effects of the molar ratio of nCu2+/nEDTA, two series of ex­
to activate PS, resulting in free Cu ion release; these released Cu ions periments were conducted. On the one hand, the Cu2+ concentration
might also participate in the PS activation. Moreover, certain Cu-based was maintained at 0.1 mmol L− 1, but the EDTA concentrations were
intermediates during Cu-EDTA decomposition might also contribute to changed to 0.1, 0.2, 0.3, and 0.4 mmol L− 1, respectively. On the other
PS activation. In this case, effective decomposition of Cu-EDTA could be hand, the EDTA concentration was maintained at 0.1 mmol L− 1, but the
achieved. However, some information on Cu-EDTA decomplexation Cu2+ concentrations were changed to 0.1, 0.2, 0.3, and 0.4 mmol L− 1,
performance and the underlying mechanisms of the PS oxidation process respectively. Other detailed experiment designs are shown in Table S1 in
triggered by Cu-EDTA itself is still unknown. the Supplementary Materials.
In this study, Cu-EDTA was selected as the target pollutant because
Cu was widely used in electroplating and electrical industries, and it is
2.3. Analysis
easily chelated with EDTA to form stable Cu-EDTA (Zeng et al., 2016a).
The removal of Cu-EDTA complex using autocatalytic PS oxidation was
Cu-EDTA was measured using high-performance liquid chromatog­
investigated, and PS activation was triggered by Cu-EDTA itself (Su
raphy (SCL-10ACP, Shimadzu) at 254 nm. 8% acetonitrile and 92%
et al., 2020). First, the influences of solution pH, initial Cu-EDTA con­
oxalic acid solutions were the eluents, and the flow rate was 1.0 mL
centration, and the molar ratio of Cu2+ and EDTA (nCu2+/nEDTA) on
min− 1. The Cu-EDTA removal processes were described using a first-
Cu-EDTA removal and PS activation were assessed. Second, the gener­
order reaction kinetic model as follows:
ation and contributions of reactive oxygen species (ROS) were evalu­
ated. Finally, the PS activation mechanisms, decomposition processes 2O•−2 + 2H + →1 O2 + H2 O2 (1)
and mechanisms of Cu-EDTA were explored via the chemical structures
of Cu-EDTA, zeta potential, intermediate generation, Cu species distri­ where C0 (mmol L− 1) represents the initial Cu-EDTA concentration and
bution, and precipitate composition. Ct (mmol L− 1) represents its residual concentration after t min of
treatment, and k (min− 1) represents the kinetic constant.
2. Experimental The solution pH was measured with a pH meter (Shanghai INESA
Scientific Instrument CO., Ltd). A manometer laser particle size meter
2.1. Chemicals (Mastersizer APA200, UK) was applied to determine the zeta potential
(ζ-potential). PS content was determined by iodimetry (Berkel et al.,
Analytically pure copper sulfate pentahydrate (CuSO4⋅5H2O, purity ˃ 2006; Liang et al., 2008). Fourier transform infrared spectroscopy
99%), iminodiacetic acid (IMDA, purity ˃ 99%), sodium persulfate (PS, (TENSOR-27, Bruker) was applied to characterize the chemical structure
purity ˃ 98%), nitrilotriacetic acid (NTA, purity ˃ 99%), ethanol (purity ˃ of Cu-EDTA. ROS was detected as reported previously (Wang et al.,
99.7%), and ethylenediaminetetraacetic acid disodium salt (EDTA-2Na, 2019a). Intermediates were detected by gas chromatograph-mass spec­
purity ˃ 99%) were provided by Sinopharm Chemical Reagent Co., Ltd., trometer (GC-MS, 7890B–5977B, Agilent, USA) and ion chromatog­
China. Chromatographically pure benzoquinone (BQ), 2,2,6,6-tetra- raphy (IC, ICS-90, DIONEX) (Wang et al., 2019a), and the pretreatment
methylpiperidine (TEMP), sodium azide (NaN3), 5,5-dimethyl-1-pyrro­ processes and detection conditions are described in S1.
line N-oxide (DMPO), and isopropanol (IPA) were provided by Sigma- After decomplexation, precipitates were collected via alkaline pre­
Aldrich. The Cu-EDTA working solution was obtained with an equal cipitation as described in S2, and the physical and chemical properties of
molar ratio of CuSO4⋅5H2O and EDTA-2Na. the precipitates were analyzed using FTIR (Vetex 70, Bruker), X-ray
diffraction (XRD, Rigaku Ultima IV, Japan), X-ray fluorescence spec­
2.2. Reaction system trometer (XRF, ARL PERFORM’X 4200, Thermo Fisher), and X-ray
photoelectron spectroscopy (XPS, Thermo ESCALAB 250XI).
Cu-EDTA removal experiments were conducted in a constant tem­
perature breeding shaker and 150 mL conical flasks. First, 100 mL of Cu- 3. Results and discussion
EDTA solution was placed in a conical flask in the shaker with stable
parameters (solution temperature T = 25 ◦ C, and rotation speed 200 r 3.1. Cu-EDTA removal under different conditions
min− 1). Under this temperature, the activation of PS by heating could be
neglected. Then, the reactions were triggered by adding 1 mL of per­ 3.1.1. PS dosage
sulfate stock solution (1 mol L− 1). After treatment, 1 mL solution was PS could be activated to produce ROS for organic pollutant removal,
collected and 100 μL pure methanol was immediately added to the so­ and its dosage significantly affects pollutant removal performance
lution to quench the reaction. This solution was used for Cu-EDTA (Fernandes et al., 2019). Fig. 1a illustrates the effects of PS dosage on
measurement. Cu-EDTA removal efficiency. The solution pH was 4, and the initial
Cu2+ stock solution with an initial concentration of 0.05 mol L− 1 was Cu-EDTA concentration was 0.3 mmol L− 1. The removal efficiency of

2
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Fig. 1. Cu-EDTA removal performance under different operation conditions (a. effect of initial PS concentration on Cu-EDTA removal; b. effect of initial Cu-EDTA
concentration on it removal; c. EPR spectra at different Cu-EDTA concentrations; d. effect of molar ratio of nCu2+/nEDTA on Cu-EDTA removal).

Cu-EDTA was 54.17% at the PS concentration of 2 mmol L− 1 after 90 exposed to a relatively lower PS concentration. Cu-EDTA removal effi­
min of treatment. This result indicated that autocatalysis of PS induced ciency was directly related to the PS dose. Due to the limited PS dosage,
by Cu-EDTA occurred, which then contributed to Cu-EDTA removal. The the Cu-EDTA removal performance was lower when the initial Cu-EDTA
Cu-EDTA removal efficiency increased with the PS concentration, concentration was higher at this stage. The following experiments were
reaching 100% when the PS concentration further increased to 80 mmol conducted with 0.1 mmol L− 1 Cu-EDTA.
L− 1 after 90 min. This could be due to the fact that the amount of ROS
increased with the PS concentration when the amount of reaction sub­ 3.1.3. Molar ratio of nCu2+/nEDTA
strates (Cu-EDTA) was constant. In addition, the consumption of radicals Fig. 1d illustrates the Cu-EDTA removal efficiency at different nCu2+/
by PS itself can also be another reason (Kan et al., 2020a). The rate nEDTA ratios. A sharp decrease in the Cu-EDTA removal was observed as
constant of parathion removal by microwave activated PS oxidation was the nCu2+/nEDTA ratio decreased from 1:1 to 1:2, and nearly no removal
also enhanced with PS dosage (Kan et al., 2020a). The concentration of performance was observed at the molar ratio of 1:4. These results
10 mmol L− 1 PS was selected for the subsequent experiments, in order to indicated that excess EDTA was not conducive to the Cu-EDTA removal.
save the use of PS. Free EDTA molecules can not only compete with Cu-EDTA for ROS, but
also chelate again with the Cu2+ released from Cu-EDTA. Excess EDTA
3.1.2. Cu-EDTA initial concentration depressed the Cu-EDTA removal in a discharge plasma oxidation system
Fig. 1b shows the Cu-EDTA removal efficiency at different initial (Wang et al., 2018). By contrast, the increase in the nCu2+/nEDTA ratio
concentrations. The Cu-EDTA removal efficiency was enhanced when promoted Cu-EDTA removal. 46.01% of Cu-EDTA was removed after 20
the initial Cu-EDTA concentration increased from 0.1 to 0.4 mmol L− 1 in min of treatment when the nCu2+/nEDTA ratio was 1:1, and the Cu-EDTA
the first 30 min. Interestingly, after 90 min of treatment, a higher Cu- removal efficiency significantly increased to 69.36% when the
EDTA removal efficiency was obtained with a lower initial concentra­ nCu2+/nEDTA ratio was 4:1. These findings suggested that excess Cu2+
tion. Typical EPR signals at different Cu-EDTA initial concentrations are participated in the PS activation, favoring Cu-EDTA removal (Wang
shown in Fig. 1c. The intensity of the ROS signals increased with et al., 2018). Lei et al. (2015) has elucidated the mechanism of PS
increasing initial Cu-EDTA concentration. These results demonstrated activation by Cu(II) and its potency for phenol removal.
that more ROS were generated in the presence of a higher concentration
of Cu-EDTA, because more amounts of PS was activated by Cu-EDTA, 3.1.4. Solution pH
which favored Cu-EDTA removal at the first treatment stage (0–30 The existing forms of organic substances in aqueous solution are
min). The species and contributions of ROS are identified and analyzed significantly affected by the solution pH value. Relatively lower solution
in Section 3.2. Wang et al. (2019c) reported that the PMS-EDTA-Fe2+ pH favored Cu-EDTA removal. Cu-EDTA removal efficiency reached up
system performed better oxidation and dewatering during sludge to 98.9% after 90 min of treatment at an initial pH of 2, and it decreased
treatment than the PMS-Fe2+ system. At the later stage (30–90 min), to 67.37% at the solution pH of 10, as shown in Fig. 2a. This might be
most of PS was consumed by Cu-EDTA, and the residual PS concentra­ due to the different hydrolysis forms of Cu-EDTA at different solution pH
tion was relatively lower when the initial Cu-EDTA concentration was values (Guan et al., 2018). As depicted in Fig. 2b, CuH2EDTA and
higher. In this case, relatively higher Cu-EDTA concentration was CuHEDTA− are the major components at the solution pH 2, CuEDTA2−

3
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Fig. 2. Effect of initial pH on Cu-EDTA removal (a. removal efficiency; b. Cu-EDTA speciation at various pH values).

and CuHEDTA− are the major components at the solution pH 4, and Cu-EDTA by the photoelectrocatalytic oxidation (Zhao et al., 2013).
CuEDTA2− is the dominant component at the solution pH values of 6, 8, Ion-exchanger/nanoscale zero-valent iron also needed 240 min to
and 10 (Zeng et al., 2016a). Completely deprotonated CuEDTA2− has remove 90% of Cu-EDTA (Liu et al., 2017). Great removal efficiencies
more difficulty reacting with ROS than the protonated Cu complexes were obtained by bicarbonate-activated H2O2 oxidation (Wang et al.,
(Wang et al., 2018). Zhao et al. (2013) also reported that basic condi­ 2019a), TiO2 photocatalysis (Lee et al., 2015), and interior micro­
tions were not beneficial for Cu-EDTA removal due to the existing form electrolysis (Ju and Hu, 2011) within 60 min. In addition, the total cost
CuOHEDTA3− . based on the equivalent removal of Cu-EDTA was also listed in Table S2.
4 and
In addition, the solution pH also affects the conversion of SO•- Herein, the total fund contained energy consumption and reagent con­
•OH in the PS oxidation system. SO•- 4 is easily converted to •OH under sumption. The total cost of this study was approximately 0.22 USD/m­
alkaline conditions. The redox potential of •OH is 2.7 V at acid condi­ mol Cu-EDTA. It was comparable with that of interior microelectrolysis
tions and 1.8 V at alkaline conditions (Zeng et al., 2016a). Therefore, (0.19 USD/mmol Cu-EDTA), but much lower than those of other
acid conditions are more conducive to the removal of pollutants in the mentioned methods. In this study, Cu-EDTA itself was used to activate
PS system. Zeng et al. (2016b) reported that acidic conditions were PS for Cu-EDTA removal, which saved the consumption of other
beneficial to the Cu-EDTA decomplexation process. chemicals. Therefore, PS oxidation can be used as an effective method to
Cu-EDTA removal performance in this study was compared with remove Cu-EDTA from wastewater.
other methods. As shown in Table S2, it took 180 min to remove 72% of

Fig. 3. ROS detection and contributions to Cu-EDTA removal (a. ROS detection by EPR; b. DMPO spin-trapping spectra; c. TEMP spin-trapping spectra; d. inhibition
rate on Cu-EDTA removal after adding ROS quenchers).

4
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

3.2. Identification of ROS for Cu-EDTA removal decreased gradually to 3.29 after 90 min of treatment, suggesting that
some small molecular acids were produced after Cu-EDTA
1
ROS including SO•- 4 , •OH, and O2 can be generated after PS acti­ decomposition.
vation (Liu et al., 2018), which would then contribute to contaminant The typical byproducts were detected and this information is shown
removal (Guo et al., 2021). EPR was used to detect the formation of ROS, in Fig. S2 and Table S3. Several byproducts, including formic acid, 2-
and its roles in Cu-EDTA removal was identified by ROS quenching hydroxypropanoic acid, oxalic acid, butanediol, ethanediol, propane­
experiments (Kan et al., 2020b; Li et al., 2020a). The signals of 1:2:2:1 diol, carbamic acid, and CH3COOH were identified. CH3COOH and NO−3
and 1:1:1:1:1:1 were observed in the EPR (Fig. 3a), which were sepa­ were gradually accumulated throughout the treatment period. Cu-IMDA
rately attributed to DMPO-•OH and DMPO-SO•- 4 , providing evidence and Cu-NTA were considered preliminary intermediates during Cu-
that SO•-
4 and •OH were generated after PS activation by Cu-EDTA. The EDTA decomposition by ROS (Wang et al., 2019a). Zhao et al. (2014)
signals of •OH were much higher than those of SO•- 4 , due to the rapid also reported that Ni-EDTA decomposed into Ni-IMDA and Ni-NTA in a
reaction of •OH with DMPO, as suggested by Wang et al. (2019c) and photoelectrocatalytic process. Then, these Cu-containing complexes
Chen et al. (2019a). As the reaction proceeded, more amounts of •OH (Cu-IMDA and Cu-NTA) were further oxidized and decomposed into
was generated, as shown in Fig. 3b. In addition, the signals of 1:1:1 were other byproducts (IMDA and NTA), accompanied by Cu release (Zhao
observed, which were attributed to TEMP-1O2, suggesting the formation et al., 2013; Huang et al., 2016; Xu et al., 2017). In addition, Toxicity
of 1O2 (Fig. 3c). Estimation Software Tool (T.E.S.T.) was applied to predict the quanti­
NaN3 and IPA were separately applied as the quenchers of 1O2 and tative structure–activity relationship of these generated byproducts, as
•OH (Huang and Zhang, 2019; Li et al., 2020b), and ethanol was applied shown in Table S4.
to capture •OH and SO•- 4 (Luo et al., 2019). Fig. S1 depicts the removal
reaction rate constants of Cu-EDTA with various quenchers. The 3.3.2. Analysis of precipitate composition
Cu-EDTA removal performance dramatically decreased after adding Chemical precipitation is widely used to remove the heavy metals
these three quenchers, suggesting that they all contributed to Cu-EDTA from water environment (Yan et al., 2019). After Cu-EDTA decom­
degardation. Then, the inhibition rate on Cu-EDTA removal after adding plexation, the released Cu species were collected by precipitation. Cu, O,
ROS quenchers was calculated based on the changes in the removal and C existed in the EDX mapping of sedimentary solids (Fig. 5), indi­
reaction rate constant (Wang et al., 2018). Fig. 3d depicts the inhibition cating the formation of Cu-based oxides and Cu-base carbonates. Char­
rate on Cu-EDTA removal after adding ROS quenchers. The inhibition acteristic peaks of CuCO3, CuO, Cu2(OH)2CO3, and Cu(OH)2 were
degrees followed the order: NaN3 ˃ ethanol ˃ IPA at the same quencher present in the XRD analysis of the precipitates (Fig. 6a). Most diffraction
dose. The inhibition rates were 96.12%, 92.29%, and 87.23% after peaks in the XRD pattern were in accordance with the XRD pattern of
adding 1.0 mmol L− 1 NaN3, ethanol, and IPA, respectively. Based on the CuO (JCPDS#80–1916), with 2θ values of 35.20, 38.42, 48.48, 53.3,
above results, it can be deduced that the roles of ROS in the Cu-EDTA 57.96, 61.24, 67.82, 72.10, and 74.88 (Abdelkader et al., 2020). This
removal process followed the order: 1O2>•OH>SO•- 4 . Wang et al. suggested that CuO was the main species in the precipitates. Similar
(2020a) also reported that SMZ decomposition was mainly attributed to results were also reported by Lee et al. (2015) and Xu et al. (2017).
4 and •OH oxidation in a Cu0–Cu2O/PS/UV system.
SO•- Fig. 6b depicts the FTIR spectrum of the precipitates. –OH vibrations
were observed at 3417 and 1612 cm− 1 (Wu et al., 2011). Cu-based
carbonates showed peaks at 1500 and 1413 cm− 1. C–O and Cu–O–O
3.3. Cu-EDTA decomposition process
vibrations were observed at 1188 and 1107 cm− 1 (Schrader et al., 2014).
Carbonates were represented at 975 and 705 cm− 1. In addition, Cu(II)-O
3.3.1. Analysis of solution composition
vibrations were observed at 619 and 543 cm− 1 (Jiang et al., 2019).
Fig. 4a depicts the ζ-potential and solution pH during the Cu-EDTA
These results suggested that some Cu-based carbonates and oxides were
removal process. The absolute value of the ζ-potential and solution pH
produced in the precipitates. The occurrence of -CO3 group in the FTIR
both decreased with the treatment time. The ζ-potential decreased from
spectrum suggested that some portions of Cu-EDTA were mineralized
− 4.82 mV to − 2.37 mV after 90 min treatment. The decrease in the
and C element in the Cu-EDTA was converted to CO2, which resulted in
absolute value of ζ-potential indicated the weaker electrostatic repulsion
the appearance of CO2− 3 in the solution.
between the organic molecules in the solution (Hu et al., 2018). In the
The valence distribution of Cu species in the precipitates was further
EDTA molecules, there existed some hydroxyl and carboxyl groups, and
explored by XPS, as shown in Fig. 7. The binding energies of approxi­
these groups could be negatively charged to produce large electrostatic
mately 934.77, 531.87, and 284.87 eV corresponding to Cu2p, O1s, and
repulsive force. After decomposition, these groups were cleaved, and the
C1s, espectively, were present (Fig. 7a). For the Cu2p, two bands at
electrostatic repulsion decreased, and the absolute value of the ζ-po­
954.47 eV and 934.47 eV were observed, which indicated that Cu(II)
tential decreased (Wang et al., 2019d). These findings demonstrated that
was the primary component of Cu species due to the energy gap of 20 eV
Cu-EDTA was decomposed by the PS oxidation. The solution pH also

Fig. 4. Variations of solution properties after Cu-EDTA decomposition (a. ζ-potential; b. solution pH).

5
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Fig. 5. EDX mapping of precipitates after treatment (a. Cu mapping; b. O mapping; c. C mapping).

Fig. 6. XRD and FTIR of precipitates after treatment (a. XRD; b. FTIR).

(Fig. 7b) (Wang et al., 2019d). For Cu2p3/2, four bands at 932.7, 933.8, 3.4. Autocatalytic activation mechanisms
934.6, and 935.0 eV corresponding to Cu(OH)2, CuO, Cu2CO3(OH)2, and
CuCO3 were separately observed (Fig. 7c) (Isahak et al., 2013), and the There were two ways for metal ions to homogeneously catalyze PS to
proportions of these peaks were 15.73%, 23.27%, 32.91%, and 28.09%, produce radicals, such as SO•-4 and •OH (Zhou et al., 2016); SO4 could be
•-

respectively. For the resolution spectrum of O1s in Fig. 7d, three peaks at easily converted to •OH, and O•- 2 could be generated by the hydrolysis
531.5, 531.2, and 529.7 eV were observed, which corresponded to and reduction of PS (Qi et al., 2016), resulting in the formation of 1O2
–C–– O, –OH, and Cu–O (Peng et al., 2016), and their proportions were and the reduction of Cu2+, as shown in reactions 2–11 (Chen et al.,
53.7%, 34.4%, and 11.9%, respectively. According to the above anal­ 2019b). Xu et al. (2017) reported that Cu-EDTA and its intermediates all
ysis, Cu(OH)2, CuO, Cu2CO3(OH)2, and CuCO3 were the main compo­ have the capability to activate PS. As shown in Fig. 8a, Cu2+, Cu-IMDA,
nents of the sediment. In addition, the precipitate composition was and Cu-NTA were selected to activate PS. Obvious signals from •OH
0.09Cu(OH)2⋅0.11CuO⋅0.079Cu2CO3(OH)2⋅0.07CuCO3 based on the were observed after the addition of certain amounts of Cu2+, Cu-IMDA,
XRF analysis. and/or Cu-NTA to the PS solution. These findings further demonstrated
that Cu2+, Cu-IMDA, and Cu-NTA can participate in PS activation. Wang
et al. (2020b) reported that the homogeneous activation of PMS by Cu2+

6
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Fig. 7. XPS of the precipitates after treatment (a. full spectrum; b. core level spectrum of Cu2p; c. resolution spectrum of Cu2p3/2; d. resolution spectrum of O1s).

Fig. 8. PS activation by Cu2+, Cu-IMDA, and Cu-NTA (a) and Cu release (b).

relied on Cu(III) and/or Cu(I). In this study, Cu(III) was not detected by the release of Cu2+. Then, PS was further activated by the Cu-containing
the periodate method (Balikungeri et al., 1977); this might be inter­ intermediates and Cu2+ via Cu(I) conversion to produce more ROS,
preted as the Cu(III) being extremely unstable, and usually not exist promoting Cu-EDTA removal. This was in accordance with the results in
under acidic conditions (Lee et al., 2013). Cu(I) was detected and its Fig. 2b, in which the Cu-EDTA removal efficiency increased with the
amount was positively related to the treatment time (Fig. 8b), indicating Cu-EDTA initial concentration in the first 30 min.
that the activation of PS by Cu2+ followed reactions 3–4 (via Cu(I)
Cu2+ + S2 O2−8 →Cu3+ + SO•−4 + SO2−4 (2)
conversion). In addition, the Cu2+ concentration was 12.5 mg L− 1 after
90 min of treatment by PS oxidation, and it decreased to 0.38 mg L− 1
Cu2+ + O•−2 →Cu+ + O2 (3)
after alkaline precipitation, which met the discharge standard of Cu (0.5
mg L− 1) of the integrated wastewater discharge standard
Cu+ + S2 O2−8 →Cu2+ + SO•−4 + SO2−4 (4)
(GB18918-2002). Based on the above analysis, the autocatalytic acti­
vation mechanisms of PS for Cu-EDTA removal were proposed. First, PS
SO•−4 + OH − →SO2−4 + • O H (5)
was activated by Cu-EDTA to produce ROS, which would attack
Cu-EDTA, resulting in the formation of Cu-containing intermediates and

7
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

SO•−4 + H2 O2 →SO2−4 + • O H + H + (6) and Cu-IMDA would also be produced after Cu-EDTA decomposition,
and then they were further decomposed to produce NTA and IMDA,
S2 O2−8 + H2 O→HO−2 + 2SO2−4 + 3H + (7) accompanied by Cu2+ release. NTA and IMDA were then oxidized and
decomposed into glycine, which was further converted to oxamic acid,
S2 O2−8 + HO−2 →SO•−4 + O•−2 + SO2−4 + H + (8) formic acid, and ammonium. Finally, CO2, H2O, and nitrate would be
generated after the decomposition of these organic acids (Xiang et al.,
Cu2+ + O•−2 →Cu+ + O2 (9) 2021).
Furthermore, the removal performances of different metal com­
O•−2 + • O H→1 O2 + OH − (10) plexes by PS oxidation, such as Cu-EDTA, Fe(III)-EDTA, and Ni-EDTA
were also evaluated. The experimental conditions were as follows. The
2O•−2 + 2H + →1 O2 + H2 O2 (11) solution temperature was 25 ◦ C, initial solution pH was 4, nMn+/nEDTA =
1:1, the PS initial concentration was 10 mmol L− 1, and the initial con­
According to the aforementioned analysis, possible decomposition
centration of each metal complex was 0.1 mmol L− 1. As shown in Fig. S3,
pathways of Cu-EDTA by the autocatalytic activation of PS are proposed
the removal efficiencies of Fe(III)-EDTA and Ni-EDTA were both lower
in Fig. 9. On the one hand, Cu-EDTA was undergone N–C bond cleavage
than 20% after 90 min of treatment. These different removal perfor­
and decomposed into CH3COOH and Cu-containing byproducts, such as
mances may be attributed to the different activation capacities of Cu-
Cu-EDDA (Xu et al., 2017). The Cu-containing byproducts were then
EDTA, Fe(III)-EDTA, and Ni-EDTA to PS. These results suggested that
oxidized to ethylene glycol and Cu2+ (Huang et al., 2016). Next,
the proposed process has a selectivity for heavy metal complexes
ethylene glycol was further oxidized to certain organic acids, which
removal.
were finally decomposed into CO2 and H2O. On the other hand, Cu-NTA

Fig. 9. Possible decomposition pathways of Cu-EDTA.

8
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

4. Conclusions Chen, L., Hu, X., Cai, T., Yang, Y., Zhao, R., Liu, C., Li, A., Jiang, C., 2019a. Degradation
of triclosan in soils by thermally activated persulfate under conditions representative
of in situ chemical oxidation (ISCO). Chem. Eng. J. 369, 344–352.
Cu-EDTA complex removal was successfully achieved by catalytic PS Chen, J., Zhou, X., Sun, P., Zhang, Y., Huang, C.H., 2019b. Complexation enhances Cu
oxidation at an acidic condition (pH = 4), and the multipathway acti­ (II)-activated peroxydisulfate: a novel activation mechanism and Cu(III)
vation of PS induced by Cu-EDTA was demonstrated. Cu-EDTA removal contribution. Environ. Sci. Technol. 53, 11774–11782.
Cuprys, A., Pulicharla, R., Lecka, J., Brar, S.K., Drogui, P., Surampalli, R.Y., 2018.
efficiency was positively related to the PS dose. PS activation was first Ciprofloxacin-metal complexes-stability and toxicity tests in the presence of humic
triggered by Cu-EDTA, and a higher removal performance was observed substances. Chemosphere 202, 549–559.
with a higher initial Cu-EDTA concentration in the early treatment stage. Dong, Y.C., Wang, P., Li, B., 2019. Fe complex immobilized on waste polypropylene
fibers for fast degradation of Reactive Red 195 via enhanced activation of persulfate
Acidic conditions and an appropriate amount of Cu2+ facilitated Cu- under LED visible irradiation. J. Clean. Prod. 208, 1347–1356.
EDTA removal; however, negative effects were observed in the pres­ Fernandes, A., Makos, P., Khan, J.A., Boczkaj, G., 2019. Pilot scale degradation study of
ence of EDTA. 1O2, •OH, and SO•- 4 were generated after PS activation,
16 selected volatile organic compounds by hydroxyl and sulfate radical based
advanced oxidation processes. J. Clean. Prod. 208, 54–64.
which contributed to Cu-EDTA removal. ROS attacks resulted in chem­ Fu, D., Kurniawan, T.A., Avtar, R., Xu, P., Othman, M.H.D., 2021. Recovering heavy
ical structure destruction of Cu-EDTA, accompanied by the formation of metals from electroplating wastewater and their conversion into Zn2Cr-layered
small molecular acids and alcohols. Cu-containing byproducts and the double hydroxide (LDH) for pyrophosphate removal from industrial wastewater.
Chemosphere 271, 129861–129874.
released Cu(II) derived from Cu-EDTA decomposition participated in PS Guan, W., Zhang, B., Tian, S., Zhao, X., 2018. The synergism between electro-Fenton and
activation, promoting more Cu-EDTA decomposition. Cu(I) conversion electrocoagulation process to remove Cu-EDTA. Appl. Catal. B Environ. 227,
was the main pathway for PS activation by Cu species. Overall, Cu- 252–257.
Guo, H., Wang, Y., Yao, X., Zhang, Y., Li, Z., Pan, S., Han, J., Xu, L., Qiao, W., Li, J.,
EDTA, as a promising autocatalytic activator, can induce the multi­
Wang, H., 2021. A comprehensive insight into plasma-catalytic removal of antibiotic
pathway activation of PS, realizing effective removal of Cu-EDTA com­ oxytetracycline based on graphene-TiO2-Fe3O4 nanocomposites. Chem. Eng. J. 425,
plex itself from wastewater. 130614–130630.
Hu, S., Hu, J., Liu, B., Wang, D., Wu, L., Xiao, K., Liang, S., Hou, H., Yang, J., 2018. In situ
generation of zero valent iron for enhanced hydroxyl radical oxidation in an
CRediT authorship contribution statement electrooxidation system for sewage sludge dewatering. Water Res. 145, 162–171.
Huang, K.Z., Zhang, H., 2019. Direct electron-transfer-based peroxymonosulfate
activation by iron-doped manganese oxide (δ-MnO2) and the development of
Qi Wang: Conceptualization, Methodology, Data treatment, Data galvanic oxidation processes (GOPs). Environ. Sci. Technol. 53, 12610–12620.
curation, Writing – original draft, Writing-Review-Editing, Writing – Huang, S., Li, Z., Chen, C., Tang, S., Cheng, X., Guo, X., 2020. Synergetic activation of
review & editing. Yutong Li: Investigation, Formal analysis, Resources. persulfate by heat and Fe(II)-complexes for hydrolyzed polyacrylamide degradation
at high pH condition: kinetics, mechanism, and application potential for filter cake
Yue Liu: Investigation, Formal analysis, Resources. Jingyu Ren: removal during cementing in CO2 storage wells. Sci. Total Environ. 713,
Conceptualization, Formal analysis, Methodology, Data treatment, Data 136561–136576.
curation. Ying Zhang: Formal analysis, Methodology. Guangzhou Qu: Huang, X., Xu, Y., Shan, C., Li, X., Zhang, W., Pan, B., 2016. Coupled Cu(II)-EDTA
degradation and Cu(II) removal from acidic wastewater by ozonation: performance,
Formal analysis, Methodology. Tiecheng Wang: Methodology, Writing-
products and pathways. Chem. Eng. J. 299, 23–29.
Review-Editing, Project administration, Funding acquisition, Writing – Isahak, W.N.R.W., Ramli, Z.A.C., Ismail, M.W., Ismail, K., Yusop, R.M., Hisham, M.W.M.,
review & editing. Yarmo, M.A., 2013. Adsorption–desorption of CO2 on different type of copper oxides
surfaces: physical and chemical attractions studies. J. CO2 Util. 2, 8–15.
Jiang, N., Qiu, C., Guo, L., Shang, K., Lu, N., Li, J., Zhang, Y., Wu, Y., 2019. Plasma-
catalytic destruction of xylene over Ag-Mn mixed oxides in a pulsed sliding discharge
Declaration of competing interest reactor. J. Hazard Mater. 369, 611–620.
Ju, F., Hu, Y.Y., 2011. Removal of EDTA-chelated copper from aqueous solution by
interior microelectrolysis. Separ. Purif. Technol. 78, 33–41.
The authors declare that they have no known competing financial Kan, H., Wang, T., Yu, J., Qu, G., Zhang, P., Jia, H., Sun, H., 2020a. Remediation of
interests or personal relationships that could have appeared to influence organophosphorus pesticide polluted soil using persulfate oxidation activated by
the work reported in this paper. microwave. J. Hazard Mater. 401, 123361–123372.
Kan, H., Wang, T., Yang, Z., Wu, R., Shen, J., Qu, G., Jia, H., 2020b. High frequency
discharge plasma induced plasticizer elimination in water: removal performance and
Acknowledgement residual toxicity. J. Hazard Mater. 383, 121185–121193.
Kim, E.J., Baek, K., 2019. Selective recovery of ferrous oxalate and removal of arsenic
and other metals from soil-washing wastewater using a reduction reaction. J. Clean.
National Natural Science Foundation of China (No. 21976143) and Prod. 221, 635–643.
Chinese Universities Scientific Fund (2452018330.7) provided financial Lan, H., Tang, Y., Zhang, X., You, S., Tang, Q., An, X., Liu, H., Qu, J., 2018.
Decomplexation of Cu(II)-EDTA over oxygen-doped g-C3N4: an available resource
supports to this study.
towards environmental sustainability. Chem. Eng. J. 345, 138–146.
Lee, H., Lee, H.J., Sedlak, D.L., Lee, C., 2013. pH-dependent reactivity of oxidants formed
Appendix A. Supplementary data by iron and copper-catalyzed decomposition of hydrogen peroxide. Chemosphere 92,
652–658.
Lee, S.S., Bai, H.W., Liu, Z.Y., Sun, D.D., 2015. Green approach for photocatalytic Cu(II)-
Supplementary data to this article can be found online at https://doi. EDTA degradation over TiO2: toward environmental sustainability. Environ. Sci.
org/10.1016/j.jclepro.2021.128119. Technol. 49, 2541–2548.
Lei, Y., Chen, C.S., Tu, Y.J., Huang, Y.H., Zhang, H., 2015. Heterogeneous degradation of
organic pollutants by persulfate activated by CuO-Fe3O4: mechanism, stability, and
References effects of pH and bicarbonate ions. Environ. Sci. Technol. 49, 6838–6845.
Li, H., Li, T., He, S., Zhou, J., Wang, T., Zhu, L., 2020a. Efficient degradation of
antibiotics by non-thermal discharge plasma: highlight the impacts of molecular
Abdelkader, T.K., Zhang, Y.L., Gaballah, E.S., Wang, S.W., Wan, Q., Fan, Q.Z., 2020.
structures and degradation pathways. Chem. Eng. J. 395, 125091–125097.
Energy and exergy analysis of a flat-plate solar air heater coated with carbon
Li, T., Wang, C., Wang, T., Zhu, L., 2020b. Highly efficient photocatalytic degradation
nanotubes and cupric oxide nanoparticles embedded in black paint. J. Clean. Prod.
toward perfluorooctanoic acid by bromine doped BiOI with high exposure of (001)
250, 119501–119511.
facet. Appl. Catal. B Environ. 268, 118442–118451.
Ahamad, T., Ruksana, Naushad, M., Al-Maswari, B.M., Alshehri, S.M., 2019. Fabrication
Liu, Y., Qu, G., Sun, Q., Jia, H., Wang, T., Zhu, L., 2021. Endogenously activated
of highly porous adsorbent derived from bio-based polymer metal complex for the
persulfate by non-thermal plasma for Cu(II)-EDTA decomplexation: synergistic effect
remediation of water pollutants. J. Clean. Prod. 208, 1317–1326.
and mechanisms. Chem. Eng. J. 406, 126774–126783.
Arruda, E.J.D., Rossi, A.P.L., Porto, K.R.D.A., Oliveira, L.C.S.D., Arakaki, A.H.,
Liu, J.J., Diao, Z.H., Liu, C.M., Jiang, D., Kong, L.J., Xu, X.R., 2018. Synergistic reduction
Scheidt, G.N., Soccol, C.R., 2010. Evaluation of toxic effects with transition metal
of copper (II) and oxidation of norfloxacin over a novel sewage sludge-derived char-
ions, EDTA, SBTI and Acrylic Polymers on Aedes aegypti (L., 1762) (Diptera:
based catalyst: performance, fate and mechanism. J. Clean. Prod. 182, 794–804.
Culicidae) and Artemia salina (Artemidae), Braz. Arch. Biol. Technol. 53, 335–341.
Liu, F., Shan, C., Zhang, X., Zhang, Y., Zhang, W., Pan, B., 2017. Enhanced removal of
Balikungeri, A., Pelletier, M., Monnier, D., 1977. Contribution to the study of the
EDTA-chelated Cu(II) by polymeric anion-exchanger supported nanoscale zero-
complexes bis(dihydrogen tellurato)cuprate(III) and argentate(III), bis(hydrogen
valent iron. J. Hazard Mater. 321, 290–298.
periodato)cuprate(III) and argentate(III). Inorg. Chim. Acta. 22, 7–14.
Berkel, K.Y., Russell, G.T., Gilbert, R.G., 2006. The dissociation rate coefficient of
persulfate in emulsion polymerization systems. Polymer 47, 4667–4675.

9
Q. Wang et al. Journal of Cleaner Production 314 (2021) 128119

Liang, C.J., Huang, C.F., Mohanty, N., Kurakalva, R.M., 2008. A rapid Wang, T., Zhou, L., Cao, Y., Zhang, Y., Zhu, L., 2019d. Decomplexation of Cu(II)-natural
spectrophotometric determination of persulfate anion in ISCO. Chemosphere 73, organic matter complex by non-thermal plasma oxidation: process and mechanisms.
1540–1543. J. Hazard Mater. 389, 121828–121837.
Luo, T., Yuan, Y., Zhou, D., Luo, L., Li, J., Wu, F., 2019. The catalytic role of nascent Cu Wang, T., Cao, Y., Qu, G., Sun, Q., Xia, T., Guo, X., Jia, H., Zhu, L., 2018. Novel Cu(II)-
(OH)2 particles in the sulfite-induced oxidation of organic contaminants. Chem. Eng. EDTA decomplexation by discharge plasma oxidation and coupled Cu removal by
J. 363, 329–336. alkaline precipitation: underneath mechanisms. Environ. Sci. Technol. 52,
Peng, B., Song, T., Wang, T., Chai, L., Yang, W., Li, X., Li, C., Wang, H., 2016. Facile 7884–7891.
synthesis of Fe3O4@Cu(OH)2 composites and their arsenic adsorption application. Wu, G., Guan, N., Li, L., 2011. Low temperature CO oxidation on Cu-Cu2O/TiO2 catalyst
Chem. Eng. J. 299, 15–22. prepared by photodeposition. Catal. Sci. Technol. 1, 601–608.
Pi, Z.J., Li, X.M., Wang, D.B., Xu, Q.X., Tao, Z.L.T., Huang, X.D., Yao, F.B., Wu, Y., He, L., Xiang, L., Xie, Z., Guo, H., Song, J., Li, D., Wang, Y., Pan, S., Lin, S., Li, Z., Han, J.,
Yang, Q., 2019. Persulfate activation by oxidation biochar supported magnetite Qiao, W., Li, J., Wang, H., 2021. Efficient removal of emerging contaminant
particles for tetracycline removal: performance and degradation pathway. J. Clean. sulfamethoxazole in water by ozone coupled with calcium peroxide: mechanism and
Prod. 235, 1103–1115. toxicity assessment. Chemosphere 283, 131156–131165.
Qi, C., Liu, X., Ma, J., Lin, C., Li, X., Zhang, H., 2016. Activation of peroxymonosulfate by Xu, Z., Shan, C., Xie, B., Liu, Y., Pan, B., 2017. Decomplexation of Cu(II)-EDTA by UV/
base: implications for the degradation of organic pollutants. Chemosphere 151, persulfate and UV/H2O2: efficiency and mechanism. Appl. Catal. B Environ. 200,
280–288. 439–447.
Schrader, I., Wittig, L., Richter, K., Vieker, H., Beyer, A., Golzhauser, A., Hartwig, A., Xu, Z., Wu, T., Cao, Y., Chen, C., Ke, S., Zeng, X., Lin, P., 2020. Efficient decomplexation
Swiderek, P., 2014. Formation and structure of copper(II) oxalate layers on carboxy- of heavy metal-EDTA complexes by Co2+/peroxymonosulfate process: the critical
terminated self-assembled monolayers. Langmuir 30, 11945–11954. role of replacement mechanism. Chem. Eng. J. 392, 123639–123649.
Shan, C., Xu, Z., Zhang, X., Xu, Y., Gao, G., Pan, B., 2018. Efficient removal of EDTA- Yan, J., Yuan, W.H., Liu, J., Ye, W.Z., Lin, J.L., Xie, J.H., Huang, X., Gao, S.S., Xie, J.H.,
complexed Cu(II) by a combined Fe(III)/UV/alkaline precipitation process: Liu, S.N., 2019. An integrated process of chemical precipitation and sulfate reduction
performance and role of Fe(II). Chemosphere 193, 1235–1242. for treatment of flue gas desulphurization wastewater from coal-fired power plant.
Su, Q., Ye, Q., Deng, L., He, Y., Cui, X., 2020. Prepared self-growth supported copper J. Clean. Prod. 228, 63–72.
catalyst by recovering Cu(II) from wastewater using geopolymer microspheres. Zeng, H., Tian, S., Liu, H., Chai, B., Zhao, X., 2016a. Photo-assisted electrolytic
J. Clean. Prod. 272, 122571–122584. decomplexation of Cu-EDTA and Cu recovery enhanced by H2O2 and electro-
Wang, B., Fu, T., An, B., Liu, Y., 2020a. UV light-assisted persulfate activation by Cu0- generated active chlorine. Chem. Eng. J. 301, 371–379.
Cu2O for the degradation of sulfamerazine. Separ. Purif. Technol. 251, Zeng, H., Liu, S., Chai, B., Cao, D., Wang, Y., Zhao, X., 2016b. Enhanced
117321–117332. photoelectrocatalytic decomplexation of Cu-EDTA and Cu recovery by persulfate
Wang, L., Xu, H., Jiang, N., Wang, Z., Jiang, J., Zhang, T., 2020b. Trace cupric species activated by UV and cathodic reduction. Environ. Sci. Technol. 50, 6459–6466.
triggered decomposition of peroxymonosulfate and degradation of organic Zhao, X., Guo, L., Hu, C., Liu, H., Qu, J., 2014. Simultaneous destruction of nickel (II)-
pollutants: Cu(III) being the primary and selective intermediate oxidant. Environ. EDTA with TiO2/Ti film anode and electrodeposition of nickel ions on the cathode.
Sci. Technol. 54, 4686–4694. Appl. Catal. B Environ. 144, 478–485.
Wang, T., Wang, Q., Soklun, H., Qu, G., Xia, T., Guo, X., Jia, H., Zhu, L., 2019a. A green Zhao, X., Guo, L., Zhang, B., Liu, H., Qu, J., 2013. Photoelectrocatalytic oxidation of Cu
strategy for simultaneous Cu(II)-EDTA decomplexation and Cu precipitation from (II)-EDTA at the TiO2 electrode and simultaneous recovery of Cu(II) by
water by bicarbonate-activated hydrogen peroxide/chemical precipitation. Chem. electrodeposition. Environ. Sci. Technol. 47, 4480–4488.
Eng. J. 370, 1298–1309. Zhao, X., Zhang, J., Qu, J., 2015. Photoelectrocatalytic oxidation of Cu-cyanides and Cu-
Wang, Q., Yu, J., Chen, X., Du, D., Wu, R., Qu, G., Guo, X., Jia, H., Wang, T., 2019b. Non- EDTA at TiO2 nanotube electrode. Electrochim. Acta 180, 129–137.
thermal plasma oxidation of Cu(II)-EDTA and simultaneous Cu(II) elimination by Zhou, P., Zhang, J., Liang, J., Zhang, Y., Liu, Y., Liu, B., 2016. Activation of persulfate/
chemical precipitation. J. Environ. Manag. 248, 109237–109245. copper by hydroxylamine via accelerating the cupric/cuprous redox couple. Water
Wang, J., Yang, M., Liu, R., Hu, C., Liu, H., Qu, J., 2019c. Anaerobically-digested sludge Sci. Technol. 73, 493–500.
conditioning by activated peroxymonosulfate: significance of EDTA chelated-Fe2+.
Water Res. 160, 454–465.

10

You might also like