You are on page 1of 10

Chemical Engineering Journal 362 (2019) 487–496

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Decomplexation of EDTA-chelated copper and removal of copper ions by T


non-thermal plasma oxidation/alkaline precipitation
Yang Caoa,b, Xuecong Qiana,b, Yuxuan Zhanga,b, Guangzhou Qua,b, Tianjiao Xiaa,b, Xuetao Guoa,b,

Hanzhong Jiaa,b, Tiecheng Wanga,b,
a
College of Natural Resources and Environment, Northwest A&F University, Yangling, Shaanxi Province 712100, PR China
b
Key Laboratory of Plant Nutrition and the Agri-environment in Northwest China, Ministry of Agriculture, Yangling, Shaanxi 712100, PR China

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• Discharge plasma/alkaline precipita-


tion is proposed to remove EDTA-Cu.
• Influences of concomitant ions on
EDTA-Cu decomplexation are eval-
uated.
• Contributions of reactive substances to
EDTA-Cu decomplexation are ana-
lyzed.
• Possible decomposition pathway of
EDTA-Cu in this system is proposed.
• Chemical compositions of precipitates
after alkaline precipitation are diag-
nosed.

A R T I C LE I N FO A B S T R A C T

Keywords: EDTA-chelated metals, which widely occur during heavy metal contaminated-soil remediation by EDTA
Non-thermal discharge plasma washing, are difficult to remove by traditional chemical precipitation because of the strong complexation be-
EDTA-chelated metals tween EDTA and heavy metal ions. A strategy, i.e., non-thermal plasma (NTP) oxidation/alkaline precipitation,
Cu-EDTA was developed to remove the EDTA-chelated metals; EDTA-chelated copper (EDTA-Cu) was used as the model
Decomplexation
pollutant and the influences of some concomitant ions on EDTA-Cu decomplexation and Cu release were eval-
Inorganic ions
uated. The concomitant anions Cl− and NO3− favored EDTA-Cu decomplexation, whereas CO32− disfavored this
process; the concomitant positive ions Ni2+ and Fe3+ both promoted EDTA-Cu decomplexation via replacement
or Fenton-like effects. These effects were also characterized by total organic carbon and Cu2+ release analysis.
The contributions of O3, %O, 1O2, %OH, and %O2− to EDTA-Cu decomplexation were quantitatively analyzed.
Ethanamine, acetamide, glycolic acid, acetic acid, formamide, ethylene glycol, and butanedioic acid were
monitored using gas chromatography-mass spectrometry, and a possible decomposition pathway of EDTA-Cu
was proposed. Furthermore, the chemical compositions of the precipitates were diagnosed by energy dispersive
X-ray spectroscopy, Raman spectroscopy, Fourier transform infrared spectroscopy, scanning electron micro-
scopy, and thermal gravimetric analysis.

1. Introduction washing is a feasible remediation method because it can permanently


eliminate heavy metals from the soil environment [1]. Various chemical
Heavy metal contamination is currently a considerable concern. Soil eluting agents were usually selected to desorb the heavy metals from


Corresponding author at: College of Natural Resources and Environment, Northwest A&F University, Yangling, Shaanxi Province 712100, PR China.
E-mail address: wangtiecheng2008@126.com (T. Wang).

https://doi.org/10.1016/j.cej.2019.01.061
Received 5 November 2018; Received in revised form 5 January 2019; Accepted 11 January 2019
Available online 14 January 2019
1385-8947/ © 2019 Elsevier B.V. All rights reserved.
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

soil particles in the soil washing process [1–4]; among which ethyle- decomposition by coexisting Cl− during the ozonation treatment pro-
nediaminetetraacetic acid (EDTA) was the most commonly used com- cess [23]. In other words, coexisting inorganic ions could exhibit dif-
plexing agent for soil remediation because of its abundant functional ferent reaction performances for pollutant elimination. However, the
groups, strong affinities with heavy metals, and minimal effects on soil influence of coexisting inorganic ions on EDTA-chelated metal elim-
properties [1,4]. Once the heavy metals were eluted with EDTA, the ination in the NTP oxidation process was still unknown.
EDTA-chelated metals in the eluent must be further treated because of In this study, EDTA-chelated metal elimination performance by the
their toxicity [5]. Traditional treatment techniques such as chemical NTP oxidation in the presence of coexisting inorganic ions was ex-
precipitation and biological treatment were ineffective for EDTA-che- plored, with EDTA-Cu as a model pollutant. The aims of this study were
lated metal removal from aqueous solution. (1) to evaluate the influences of coexisting inorganic ions, including
Recently, a two-step strategy, preoxidation for the EDTA-chelated Cl−, CO32−, NO3−, Ni2+, and Fe3+, on EDTA-Cu decomplexation and
metals and post-precipitation for free heavy metal ions, has become an copper ion release; (2) to identify the detailed contributions of the re-
attractive alternative for EDTA-chelated metal elimination from aqu- active substances to EDTA-Cu decomplexation; and (3) to further ex-
eous solution [6–9]. The EDTA-chelated metals were first oxidized and plore the mineralization and precipitation process.
decomposed by oxidation methods, accompanied by the release of free
heavy metal ions into solutions; then, post-precipitation was used to 2. Experimental
further precipitate these heavy metal ions. Recent research has reported
that the EDTA-chelated metals could be oxidized and decomposed by 2.1. Experimental chemicals
photo-Fenton [3], photocatalysis [6], photoelectrocatalytic oxidation
[8], ultraviolet (UV)/H2O2 treatment [9], and ozonation [7], which Analytical purity EDTA-2Na (purity > 99.0%) and CuSO4
then resulted in the release of free metal ions. However, acidic condi- (purity > 99.95%) were bought from the Sinopharm Chemical Factory
tions were necessary for Fenton oxidation, and iron sludge was also in China. Analytical-grade NaNO3, Na2CO3, NaCl, Ni(NO3)2 and Fe
needed for further treatment. Potential risks were posed by the utili- (NO3)3 were all bought from the Tianjin Fine Chemical Factory in
zation of nano-TiO2 in photocatalytic or photoelectrocatalytic oxida- China. The stock solution of EDTA-chelated copper with an initial
tion. H2O2 oxidation alone could not realize the EDTA-chelated metal concentration of 50 mmol L−1 was prepared by mixing EDTA-2Na so-
decomplexation, and must generally be combined with other techni- lution and CuSO4 solution, and their molar ratio was 1:1. Prior to use,
ques, which inevitably increased the operating cost. The direct oxida- the working solution was obtained by diluting the stock solution to
tion of ozone was generally poor and it had a high selectivity for or- 0.3 mmol L−1. 1,4-Diazabicyclooctane triethylenediamine (DABCO,
ganic contaminants; thus catalytic oxidation of ozone was necessary. purity > 99.0%), isopropanol (IPA, purity > 99.8%), and benzoqui-
Therefore, it was necessary to develop new approaches for the elim- none (BQ, purity > 99.5%) were brought from Sigma-Aldrich. All
ination of EDTA-chelated metals from aqueous solution. other chemicals were analytically pure and used without further pur-
Non-thermal plasma (NTP) was one kind of advanced oxidation ification.
techniques which has attracted lots of attention recently in environ-
mental pollution control [10–14]. Compared to other classical methods, 2.2. NTP reaction system
NTP has been proven to possess multiple advantages for pollution
control, such as high efficiency, a short time-consuming, operation at The NTP reaction system for EDTA-chelated copper removal is
normal temperature and pressure, environmental compatibility, and no shown in Fig. 1. Specifically, an alternating current high-voltage power
exogenous chemical utilization [15–18]. In the NTP process, electrons supply (discharge voltage 0–30 kV and frequency 50 Hz) was employed
could be accelerated and gained a high temperature of 104–105 K when to provide high voltage and form discharge plasma. This power supply
a certain voltage was applied across the electrodes, whereas the back- was purchased from Dalian University of Technology, China. The re-
ground gas still maintained at a normal temperature. The high-energy action vessel was made of Plexiglass cylinder. High voltage electrode
electrons then collide with background molecules to generate a large was made of a stainless steel spring with specifications of 1.0 mm dia-
number of reactive substances, such as %OH, %O, %O2−, 1O2, and HO2% meter and 18 cm length. Insulation dielectric was made of a quartz glass
[15–18]. These reactive substances can oxidize most organic con- cylinder (300 mm height and 1.0 mm wall thickness), and an aerator
taminants. Other long-lived molecules, such as O3 and H2O2, were also was connected to its bottom. The stainless steel spring was tightly fitted
produced and played significant roles in pollution control. Previous into the quartz glass cylinder. Dry air was used as the carrier gas. When
studies have mainly focused on organic pollutant removal using NTP, a certain discharge voltage was applied across the electrodes, the dry
while only few studies were carried out on free heavy metal removal
[19,20]. In our previous study, a surface discharge plasma reaction
system was designed to remove complexing agent (humic acid) from
water. In this reaction system, plasma was first triggered in the gas
atmosphere and reactive substances were quickly poured into liquid
phase in the form of small bubbles by gas-flow. And thus solution
conductivity did not affect the reaction process and the utilization ef-
ficiency of the reactive substances would be also promoted [10].
Elimination of EDTA-chelated copper was effectively realized by the
surface discharge plasma system in a short treatment time, accom-
panied by high-efficient mineralization [21]. However, the detailed
contributions of the reactive substances to EDTA-chelated copper re-
moval were still not identified.
Inorganic ions, such as Cl−, SO42−, CO32−, NO3−, and Fe3+,
usually existed in the soil environment and were also eluted in the soil
washing process. Previous studies reported that these inorganic ions
possibly participate in oxidation processes [22–24]. For example,
CO32− could favor dye-containing wastewater decoloration, while Cl−
and SO42− disfavored decoloration in an ozonation process [24]. Fur-
thermore, Acid Red 27 decoloration was promoted due to ozone Fig. 1. Schematic of the experimental setup for EDTA-Cu decomplexation.

488
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

air would be excited and ionized to form NTP along the quartz glass
tube, and then the generated reactive substances were poured into the
solution. 500 mL 0.3 mmol L−1 EDTA-Cu wastewater was treated in a
batch processing and the voltage was 16 kV unless otherwise specified.

2.3. Methods and analysis

After the EDTA-Cu was decomposed by the NTP, the solution pH


value was adjusted to 11.5 using NaOH solution to precipitate the re-
leased free copper ions. After natural settlement for 48 h, the pre-
cipitates were collected by vacuum filtration. The obtained solid sam-
ples were characterized by Fourier transform infrared spectroscopy
(FTIR, Nicolet iS10, Thermo), scanning electron microscopy (SEM, S-
3400N, Hitachi), energy dispersive X-ray (EDX, Czech Republic,
TESCAN), Raman spectroscopy (France, LabRAM HR800), and thermal
gravimetric analyzer (TGA, Germany, TG209C). Atomic absorption
spectrophotometer (ZA3000, Hitachi) was used to measure the copper
concentration in liquid. Cu removal efficiency, EDTA-Cu decomplexa-
tion kinetics, and energy efficiency (EE) were calculated as previous
Fig. 2. Effect of Cl− on EDTA-Cu decomplexation (a. decomplexation effi-
study [21]. Inhibition rate on EDTA-Cu decomplexation in the presence
ciency; b. decomplexation kinetics).
of a reactive substance trapping agent was calculated according to the
following equation,
constant was observed at a Cl− concentration of 0.5 mmol L−1, as de-
k − kad picted in Fig. 2b.
Inhibition rate (%) = × 100
k (1) Previous study reported that Cu2+ in solution could chelate with

where k and kad are the first-order kinetic rate constants without and Cl to form complex ions, as shown in reaction (2) [25]. Therefore, a
with the trapping agent, respectively. low concentration of Cl− might compete with EDTA molecules for
A fluorospectro photometer (LS55, PerkinElmer) was employed to Cu2+, which then favored EDTA-Cu decomplexation. On the other
characterize %OH radical. The EDTA-Cu concentration was determined hand, Cl− could also consume reactive substances such as %OH radical,
using a HPLC as reported in previous study [21]. The total oxidizing ozone, and H2O2, resulting in their invalid loss, as shown in reactions
capacity was measured using iodometry. The combination of heavy (3)–(5) [26,27]. These reactive substances were reported to play vital
metals with EDTA was characterized by an UV–Vis spectrophotometer roles in pollutant decomposition [12,13]. However, it must be noted
(ND-2000c, Thermo). The generated intermediates after EDTA-chelated that Cl%, HClO, and HO2% could be generated via the reactions of Cl−
copper decomplexation were detected using GC–MS (7890B-5977B, with %OH, ozone, and H2O2, which were all strong oxidants and might
Agilent). Prior to GC–MS detection, the extraction and pretreatment partly compensate for the loss of %OH, ozone, and H2O2. In summary,
procedures were as follows: (1) the solution was extracted using ethyl the dose-effect should be taken into consideration for the influence of
acetate several times, and the extracting solution was then concentrated Cl−. Ramjaun et al. found that the presence of Cl− disfavored Reactive
to near-dry in a rotary evaporator; (2) subsequently, 20 μL pyridine, Red decomposition by gliding arc plasma oxidation [26]. Merouani
200 μL dioxane, and 100 μL silane reagent (trimethylchlorosilane and et al. also reported that dyes/dye mixture decoloration performance by
trifluoroacetamide mixture) were added to the above concentrated gliding arc plasma oxidation decreased after Cl− addition [28]. The
sample; (3) finally, the mixture was derivatized for 20 min at 60 °C. For changes in the relative content of %OH at various Cl− concentration
GC–MS analysis, the column temperature was set at 50 °C for 5 min, addition are shown in Fig. 2a (inset). The relative content of %OH de-
increased to 100 °C at a rate of 5 °C/min, and was held at 100 °C for creased at a higher Cl− concentration, and there was an approximately
1 min. Subsequently, it rose to 250 °C at a rate of 10 °C/min and kept for 53.5% decline after 5 min of treatment with 1.0 mmol L−1 Cl− addi-
1 min. The MS was operated using electron-impact mode and scanned tion. These results further confirmed that concomitant Cl− would
in the m/z range of 30–500. Helium was used as the carrier gas. The consume some reactive substances.
source and quadrupole temperatures were held at 230 °C and 150 °C, Cl− + Cu2+ → [CuCl 4 ]2 − (2)
respectively.

3. Results and discussion Cl− + %OH → %Cl + OH− (3)


Cl− + O3 + H+ → HClO + O2 (4)
3.1. EDTA-Cu decomplexation performance in the presence of anions

3.1.1. Influence of Cl− Cl% + H2O2 → HO2% + Cl− + H+ (5)


Fig. 2a depicts the evolution of EDTA-Cu decomplexation efficiency
with treatment time at various Cl− concentrations. As a control ex-
periment, EDTA-Cu decomplexation in the presence of Cl− was also 3.1.2. Influence of CO32−
conducted when the NTP system was off. No reactive substances were Fig. 3a depicts the evolution of EDTA-Cu decomplexation efficiency
formed and no displacement reactions occur. Therefore, EDTA-Cu de- with treatment time at various CO32− concentrations. EDTA-Cu de-
complexation in this case was negligible. The presence of Cl− promoted complexation in the presence of CO32− was also conducted when the
EDTA-Cu decomplexation in the NTP system. Approximately 71.7% of NTP system was off. No reactive substances were formed and no dis-
EDTA-Cu was decomposed within 30 min of treatment without Cl− placement reactions occur. Therefore, the EDTA-Cu decomplexation in
addition, whereas it increased to 88.4% when the Cl− concentration this case was negligible. The presence of CO32− disfavored EDTA-Cu
was 0.5 mmol L−1. And further increasing the Cl− concentration to decomplexation in the NTP system. The EDTA-Cu decomplexation ef-
1.0 mmol L−1, the EDTA-Cu decomplexation efficiency reached 81.8% ficiency decreased by 14.4% within 30 min of treatment when the
within 30 min of treatment. Accordingly, the highest reaction kinetic CO32− concentration was 0.5 mmol L−1, compared with that without

489
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

decomplexation in the presence of NO3− was also conducted when the


NTP system was off. No reactive substances were formed and no dis-
placement reactions occur. Therefore, EDTA-Cu decomplexation in this
case was negligible. The presence of NO3− promoted EDTA-Cu de-
complexation to a certain extent in the NTP system. The EDTA-Cu de-
complexation efficiency increased by 21.6% within 20 min of treatment
when the NO3− concentration was 0.1 mmol L−1 compared with that
without NO3− addition. And further increasing the NO3− concentration
to 1.0 mmol L−1, the EDTA-Cu decomplexation efficiency then de-
creased to 64.2%. Accordingly, the highest reaction kinetic constant
was observed with the addition of 0.1 mmol L−1 NO3− (Fig. 4b).
Previous study reported that NO3− could be converted into O and %
OH under ultraviolet irradiation, as shown in reactions (7) and (8) [32].
Therefore, the promoted EDTA-Cu decomplexation with the addition of
a low concentration of NO3− was probably attributed to the enhanced
formation of reactive substances such as O and %OH because ultraviolet
irradiation was produced concomitantly in the NTP system. On the
other hand, NO3− would probably consume %OH because their reaction
rate was approximately 1.4 × 108 M−1 s−1 (reaction (9)) [32,33]. This
Fig. 3. Influence of CO32− on EDTA-Cu decomplexation (a. decomplexation
reaction would lead to NO3% radical formation, which has an oxidizing
efficiency; b. decomplexation kinetics).
ability of 2.30–2.60 V [33]. Therefore, high-efficient EDTA-Cu decom-
plexation efficiency was observed with 1.0 mmol L−1 NO3− addition
CO32− addition. Correspondingly, the reaction kinetics also gradually due to compensation by the NO3% radical. This deduction was further
decreased (Fig. 3b). confirmed by the changes in the relative content of %OH with different
It was reported that CO32− could act as a trapping agent of %OH due amounts of NO3− addition (Fig. 4a). The highest %OH content was
to the rapid reaction (approximately 4.0 × 108 M−1 s−1) [29,30]. This detected with 0.1 mmol L−1 NO3− addition.
could further be confirmed by the changes in the fluorescence signal of
NO−3 + hv → O+ NO−2 (7)
%OH at different amounts of CO32−. The relative content of %OH radical
decreased gradually with the increase of CO32− concentration, as de-
picted in Fig. 3a. Furthermore, the presence of CO32− would provide O + H2O → 2%OH (8)
alkaline conditions for EDTA-Cu decomplexation. Previous study has

confirmed that alkaline conditions were not beneficial for EDTA-Cu NO3 + %OH → H2O + NO3% (9)
decomplexation in a discharge plasma system [21]. Therefore, an in-
hibitory effect for EDTA-Cu decomplexation was observed in the pre-
3.2. EDTA-Cu decomplexation performance in the presence of positive ions
sence of CO32− in this study. Huang et al. also found that there was an
approximately 49% decline in methylene blue decoloration efficiency
3.2.1. Influence of Ni2+
in a dielectric barrier discharge plasma system with 7.5 mmol L−1
Fig. 5a depicts the evolution of EDTA-Cu decomplexation efficiency
CO32− addition [31].
with treatment time at various Ni2+ concentrations. The presence of
CO32− + %OH → H2O + CO3−% (6) Ni2+ significantly promoted EDTA-Cu decomplexation in the NTP
system. The EDTA-Cu decomplexation efficiency was only 71.7%
within 30 min of treatment in the NTP system without any Ni2+ addi-
3.1.3. Influence of NO3− tion; whereas it was gradually improved to 90.5%, 92.5%, and 94.4%
Fig. 4a depicts the evolution of EDTA-Cu decomplexation efficiency with 0.1, 0.5, and 1.0 mmol L−1 Ni2+ addition, respectively. Corre-
with treatment time at various NO3− concentrations. EDTA-Cu spondingly, the reaction kinetic constant also increased gradually with

Fig. 4. Evolution of EDTA-Cu decomplexation with treatment time at various Fig. 5. Influence of Ni2+ on EDTA-Cu decomplexation (a. decomplexation ef-
NO3− additions (a. decomplexation efficiency; b. decomplexation kinetics). ficiency; b. decomplexation kinetics).

490
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

Ni2+ concentration, as shown in Fig. 5b.


%OH radical was probably generated via the reaction of Ni2+ with
H2O2 (reaction (10)) [34], which then favored pollutant decomposition.
To explore this point, the changes in the relative content of %OH with
different amounts of Ni2+ addition were evaluated, and the results are
shown in Fig. 5a. No significant change in %OH content was observed
with the addition of different amounts of Ni2+. This result suggested
that the reaction between Ni2+ and H2O2 did not occur in this study.

Ni2+ + H2O2 → Ni3+ + OH− + %OH (10)

Ni2 + + EDTA - Cu ↔ Cu2++EDTA - Ni (11)

EDTA molecules could chelate with lots of metal ions with different
stability constants. The stability constant (lgK) of EDTA-Cu was ap-
proximately 18.80, which was comparable to that of EDTA-Ni
(lgK = 18.67) [35]. Therefore, Ni2+ in solutions might compete with
Cu2+ for EDTA molecules, resulting in EDTA-Cu decomplexation (see
reaction (11)). This competing reaction was confirmed via the evolution
of UV–Vis spectra and HPLC graphs of EDTA-Cu/Ni solutions when the Fig. 7. Influence of Fe3+ on EDTA-Cu decomplexation (a. decomplexation ef-
NTP was off. The typical UV–Vis spectra of EDTA-Cu significantly ficiency; b. decomplexation kinetics).
changed in the presence of Ni2+, as shown in Fig. 6a. More obviously,
the EDTA-Cu signal in the HPLC graph gradually decreased with time
3.2.2. Influence of Fe3+
continued when a certain amount of Ni2+ was added (the NTP was not
Fig. 7a depicts the evolution of EDTA-Cu decomplexation efficiency
triggered; and EDTA-Cu concentration decreased by approximately
with treatment time at various Fe3+ concentrations. The presence of
23% within 30 min with 1.0 mmol L−1 Ni2+ addition), whereas the
Fe3+ significantly promoted EDTA-Cu decomplexation in the NTP
EDTA-Ni signal increased, as depicted in Fig. 6b. Thus, the promoted
system. The EDTA-Cu decomplexation efficiency increased gradually
EDTA-Cu decomplexation efficiency could be attributed to the re-
with the Fe3+ concentration, and there was an approximately 52.9%
placement effect of Ni2+.
rise in the EDTA-Cu decomplexation efficiency within 10 min of treat-
ment when the Fe3+ concentration was 1.0 mmol L−1 compared with
that without Fe3+ addition. Correspondingly, the highest reaction ki-
netic constant was observed with 1.0 mmol L−1 Fe3+ addition, as
shown in Fig. 7b.
The promoted EDTA-Cu decomplexation in the presence of Fe3+
could partly be attributed to the Fenton-like effect of Fe3+/H2O2. %OH
radicals could generate via the reaction of Fe3+ with H2O2 (see reac-
tions (12) and (13)) [36]. This deduction was confirmed by the changes
in the relative content of %OH with different amounts of Fe3+ addition,
as shown in Fig. 7a. There was an approximately 40.7% rise in %OH
content with 1.0 mmol L−1 Fe3+ addition. On the other hand, the sta-
bility constant (lgK) of EDTA-Fe(III) was approximately 25.10, which
was larger than that of EDTA-Cu (lgK = 18.80) [35]. That is, Fe3+ in
solution easily competed with Cu2+ for EDTA molecules, resulting in
EDTA-Cu decomplexation (see reaction (14)). This competing effect
was clearly observed from the evolution of the UV–Vis spectra and
HPLC graphs of the EDTA-Cu/Fe solutions when the NTP was off. The
typical UV–Vis spectrum of EDTA-Cu significantly changed in the pre-
sence of Fe3+, as shown in Fig. 8a. The EDTA-Cu signal in the HPLC
graph also decreased with time continued when a certain amount of
Fe3+ was added (the NTP was not triggered, and EDTA-Cu concentra-
tion decreased by approximately 31.1% within 10 min with
1.0 mmol L−1 Fe3+ addition), whereas the EDTA-Fe(III) signal in-
creased, as depicted in Fig. 8b. Therefore, the replacement effect of
Fe3+ and Fenton-like effect jointly promoted EDTA-Cu decomplexation
in the NTP system.

Fe3+ + H2O2 → Fe2+ + H+ + HO2% (12)

HO2% + H2O2 → O2 + H2O + %OH (13)


Fe3 + + EDTA - Cu → Cu2 ++EDTA - Fe (14)

3.3. EDTA-Cu decomplexation process

To further explore the EDTA-Cu decomplexation process, TOC and


Fig. 6. Changes in UV–Vis and HPLC of EDTA-Cu samples with/without Ni2+ Cu removal in the presence of ions were evaluated, and the results are
(a. UV–Vis spectrum; b. HPLC chromatograph). manifested in Fig. 9. Similar to the decomplexation efficiency, the

491
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

Fig. 10. Effects of reactive substance trapping agents on EDTA-Cu decom-


plexation.

reactions (15)–(33) [14,21,37]. To distinguish their contributions to


EDTA-chelated copper decomplexation, the influences of reactive sub-
stance trapping agents on EDTA-Cu decomplexation were evaluated.
1,4-Diazabicyclooctane triethylenediamine (DABCO), isopropanol
(IPA), and benzoquinone (BQ) were generally chosen as trapping agents
for 1O2, %OH, and %O2−, respectively, because their each reaction rate
was about 2.0 × 109 [38–40]. Fig. 10 depicts the evolution of EDTA-Cu
decomplexation efficiency with treatment time with the addition of
different trapping agents. The content of each trapping agent was
0.5 mmol L−1, which inhibited EDTA-Cu decomplexation to the
greatest extent. The EDTA-Cu decomplexation efficiency decreased to
50.3%, 41.7%, 28.4%, 32.4%, 15.4%, 14.3%, and 9.4% after 50 min of
Fig. 8. Changes in UV–Vis and HPLC of EDTA-Cu samples with/without Fe3+ treatment with IPA, DABCO, BQ, “IPA + DABCO”, “BQ + IPA”,
(a. UV–Vis spectrum; b. HPLC chromatograph). “BQ + DABCO”, and “DABCO + IPA + BQ” addition, respectively,
suggesting that the reactive substances including 1O2, %OH, and %O2−
all participated in the EDTA-chelated copper decomplexation, espe-
cially %O2− species. In our previous study, electron paramagnetic re-
sonance was used to detect the formation and changes of %OH, 1O2, and
%O2− in the NTP system; in which the signal intensities of %OH, 1O2,
and %O2− were lower in the presence of EDTA-Cu than those without
EDTA-Cu, but they all increased as discharge voltage [21]. These results
could also confirm the roles of %OH, 1O2, and O2%−.
Based on the EDTA-Cu decomplexation kinetics, the inhibition rate
of EDTA-Cu decomplexation was calculated and presented in Table 1.
The inhibition rate reached 74.5%, 87.6%, and 80.4% with solely IPA,
BQ, and DABCO addition, respectively; thus, the total contribution of %
OH, %O2−, and 1O2 to EDTA-Cu decomplexation was 74.5%, 87.6%,
and 80.4%, respectively. For “IPA + BQ + DABCO” addition, %OH, %
O2−, and 1O2 were consumed by the corresponding trapping agents,
which led to the highest inhibition rate on EDTA-Cu decomplexation
(95.9%); therefore, the direct contribution of O3 and %O to EDTA-Cu
Fig. 9. TOC and Cu removal after EDTA-Cu decomplexation at various ions decomplexation was 4.1% (100% − 95.9% = 4.1%).
addition. The inhibition rate was 85.1% for “IPA + DABCO” addition, and O3,
%O, and %O2− contributed to EDTA-Cu decomplexation. Some %O and
presence of NO3−, Cl−, Ni2+, and Fe3+ favored TOC and Cu removal, O3 was converted to %OH and 1O2 (reactions (21), (23), and (24)), but
whereas the presence of CO32− inhibited this process. The highest TOC they could only direct participate in EDTA-Cu decomplexation. There-
and Cu removal efficiencies reached 86.7% and 88.9% within 50 min of fore, the contribution of %O2− derived from O3 reactions (reactions (25)
treatment in the presence of Fe3+. and (26)) was 10.8% (100% − 85.1% − 4.1% = 10.8%); and other
pathways leading to %O2− formation occupied 76.8%
3.4. Identification of reactive substance responsibility (87.6% − 10.8% = 76.8%) of the contribution to EDTA-Cu decom-
plexation.
The NTP was first triggered in the gaseous environment, and the Approximately 93.3% of inhibition rate was observed for
generated gaseous plasma was then transferred to the liquid environ- “IPA + BQ” addition, and O3, %O, and 1O2 contributed to EDTA-Cu
ment in this study. In this process, some oxidative reactive substances, decomplexation. Some %O and O3 was converted to %OH and %O2−
including O3, %O, 1O2, %OH, and %O2−, would generate, as shown in (reactions (21)–(26), and (30)), but they could only direct participate in

492
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

Table 1
Reaction rate (k) and inhibition rate with different trapping agents.
Trapping agents k (min−1) Quenching substances Detection of substances Inhibition rate (%)

No 0.051 – O3, %O,%OH, O2, %O2−


1
0
IPA 0.013 %OH O3, %O,1
O2, %O2− 74.5
BQ 0.0063 %O2− O3, %O,%OH, 1O2 87.6
DABCO 0.01 1
O2 O3, %O,%OH, %O2− 80.4
IPA + DABCO 0.0076 %OH, 1O2 O3, %O,%O2− 85.1
IPA + BQ 0.0034 %OH, %O2− O3, %O,1
O2 93.3
BQ + DABCO 0.0032 1
O2, %O2− O3, %O, %OH 93.7
IPA + BQ + DABCO 0.0021 %OH, 1O2, %O2− O3, %O 95.9

%O + O2 + M → O3 + M (20)

%O + H2O → %OH (21)

%O + H2O → H2O2 (22)

H2O2 + O3 → HO2− + O2 + %OH (23)

O3 + H2O → 2%OH + O2 (24)



O3 + OH → HO2% + %O2− (25)

O3 + %OH → %O2 + H +
(26)

%O2 + H → HO2%
+
(27)

HO2% + H + e → H2O2
+
(28)

O3 + NO2− → 1O2 + NO3− (29)


− −
%O2 + H2O2 → O2 + OH + %OH
1
(30)
Fig. 11. GC–MS chromatograph of the intermediates during EDTA-chelated
− −
copper decomplexation after NTP treatment. %O2 + %OH → O2 + OH
1
(31)

%O2− + HO2% → 1O2 + OH2− (32)


EDTA-Cu decomplexation. Thus, 1O2 rooted from %O and O3 reaction
(reaction (29)) contributed to 2.6% of EDTA-Cu decomplexation HO2% + HO2% → 1O2 + H2O2 (33)
(100% − 93.3% − 4.1% = 2.6%), and other ways resulting in 1O2
generation accounted for 77.8% (80.4% − 2.6% = 77.8%) of the con-
tribution. 3.5. Identification of byproducts during EDTA-Cu decomplexation
The inhibition rate was 93.7% for “BQ + DABCO” addition, and O3,
%O, and %OH contributed to EDTA-Cu decomplexation. Some %O and O3 GC–MS was applied to further identify the byproducts of EDTA-
was converted to %O2− and 1O2 (reactions (25), (26), and (29)–(33)), chelated copper after 50 min of treatment in the NTP system, and the
but they could only direct participate in EDTA-Cu decomplexation. total chromatograph and detailed structural formulas of byproducts are
Thus, %OH rooted from %O and O3 reactions (reactions (21), (23) and depicted in Fig. 11 and Table 2. Some silicane derivatives were
(24)) contributed to 2.2% (100% − 93.7% − 4.1% = 2.2%) of EDTA-
Cu decomplexation, and other ways resulting in %OH generation ac- Table 2
counted for 72.3% (74.5% − 2.2% = 72.3%) of the contribution. Retention times and structural formulas of byproducts.
As discussed above, %O and ozone formed in the gaseous environ-
Serial Retention time Byproducts m/z Structural formula
ment in the NTP system could not only directly participate in reactions
number (min)
related to EDTA-Cu decomplexation but could also lead to the forma-
tion of other reactive substances, including %OH, 1O2, and %O2−, which 1 10.338 Ethanamine 189
then contributed to EDTA-Cu decomplexation. However, it should be
noted that the direct contributions of %O and ozone were quite low, and
their roles were mainly indirect (formation of other reactive sub- 2 11.143 Acetamide 131
stances). It was difficult to quantificationally distinguish the con-
centrations of these reactive substances. Iodometry was used to eval-
uate the total oxidizing capacity of the NTP system, which was 3 11.498 Formamide 189

characterized as ozone equivalent concentration. The concentration


reached 3.5 mg L−1 within 10 min of NTP treatment.
4 12.064 Ethylene glycol 191
N2 + e → N(4S) + N(2D) + e (15)
5 14.668 Glycolic acid 205
O2 + N(2D) → %NO + O (16)

%NO + %OH → NO2− + H+ (17) 6 19.308 Acetic acid 132

O2 + e → e + 2%O (18)
7 20.361 Butanedioic acid 262
O2 + e → %O2− (19)

493
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

Fig. 12. Possible decomplexation pathway of EDTA-chelated copper in the NTP


system.

detected, and their matrixes included ethanamine, acetamide, for- Fig. 14. SEM photo of the precipitates after 50 min treatment.
mamide, ethylene glycol, glycolic acid, acetic acid, and butanedioic
acid. These results indicated that the above matrix byproducts were subsequently converted to acetic acid and acetamide; and acetamide
generated during EDTA-chelated copper decomplexation in the NTP could also be oxidized to formamide and formic acid. The generated
system. Iminodiacetic acid (IMDA), glycine, glycolic acid, oxamic acid, glycolic acid could be further oxidized to form butanedioic acid and
CH3COOH, oxalic acid, and HCOOH were detected during EDTA-che- oxalic acid, which would finally be converted into formic acid. On the
lated nickel decomplexation by photoelectrochemical oxidation [41]. other hand, CeN bonds in position (2) could also be attacked by re-
Zhao et al. reported that EDA, glycine, oxamic acid, oxalic acid, acetic active substances to generate ethylene glycol and IMDA. Then ethylene
acid, and formic acid would generate during EDTA-chelated copper glycol was probably converted to glycolic acid, which finally would
decomplexation by photoelectrocatalytic oxidation [8]. EDA, glycine, lead to formic acid formation. The generated IMDA was possibly oxi-
glyoxylic acid, oxamic acid, and oxalic acid were also detected during dized to glycine, which was then conducive for glycolic acid and
EDTA-chelated copper decomplexation by ozonation [42]. Oviedo et al. oxamic acid formation. The oxamic acid was easily oxidized to form
found that glycine, butanedioic acid, acetic acid, and formic acid were oxalic acid. Certainly, the decomplexation process of EDTA-chelated
generated during Fe(III)-EDTA abatement by Fenton oxidation [43]. copper would lead to copper ion release [44–46].
Based on the identified byproducts and available references, pos-
sible decomplexation pathways of EDTA-chelated copper in the NTP
3.6. Precipitate composition analysis
system are proposed in Fig. 12. On the one hand, CeN bonds in position
(1) could be attacked by reactive substances to form EDA and glycolic
To identify the microstructure and constituents of the Cu-containing
acid. Then, EDA would be further oxidized to ethanamine, which was
precipitates, EDX mapping, SEM, FTIR, Raman, and TG-DTG were

Fig. 13. EDX analysis of the precipitates after 50 min of treatment (a. EDX total mapping; b. C mapping of EDX; c. O mapping of EDX; d. Cu mapping of EDX).

494
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

mode was observed at approximately 1610 cm−1 [50], suggesting that


Cu-based hydroxides such as Cu(OH)2 and Cu2CO3(OH)2 might also be
generated. Cu(OH)2 and Cu2CO3(OH)2 were also detected by Huang
et al. after EDTA-chelated copper was decomposed by ozonation [42].
Xu et al. reported that copper ions in the precipitates mainly existed as
Cu2+, which could bind to oxygen to form CuO after EDTA-chelated
copper was eliminated by UV/H2O2 or UV/persulfate [30]. Cu2+ was
also efficiently formed during EDTA-chelated copper decomplexation
by photoelectrocatalytic oxidation, and then it would be converted to
Cu0 onto the cathode via electroreduction [8]. However, Cu+ or Cu0
did not exist in the strongly oxidized environment in the NTP system.
Fig. 15(b) depicts the Raman spectrum of the precipitates. The ab-
sorption bands at 276, 324, and 606 cm−1 corresponded to the Raman
shifts of CuO and the peak at 495 cm−1 could be associated with the
Raman shift of Cu(OH)2 [51]. In addition, the absorption bands at
1046, 1363, and 1521 cm−1 corresponded to the symmetric vibration
of CO32− [52]. These results further confirmed that copper oxides, Cu
(OH)2, and other Cu-based carbonates were generated in the pre-
cipitates.
The weight loading of the precipitates was analyzed using thermo-
gravimetric (TG) and derivative thermogravimetric (DTG) analysis. The
TG spectrum showed a 4.84% weight loss from 25 °C to 205 °C, and the
highest thermal oxidation rate was observed at 200 °C, as depicted by
the DTG spectrum; this thermal distortion could be attributed to the
dehydration of Cu(OH)2, as shown in Fig. 15(c) [53]. There was a
4.31% weight loss from 205 °C to 345 °C with the highest thermal
oxidation rate of 239 °C, as shown in the TG and DTG spectra; which
could be assigned to Cu2(OH)2CO3 decomposition [54]. 1.63% weight
loss from 345 °C to 495 °C with the highest thermal oxidation rate of
376 °C, and 13.01% weight loss from 495 °C to 900 °C with the highest
thermal oxidation rate of 812 °C could be associated with CuCO3 and
CuO decomposition, respectively [54,55]. Frost et al. reported that
(Zn2,Cu3)(CO3)2(OH)6 was decomposed to form ZnCO3 and CuCO3 at
235–280 °C, ZnCO3 was decomposed to ZnO at 394 °C, and CuO could
be decomposed at 805 °C, as analyzed by TG-DTG [54]. Tsai et al. also
found that little mass loss of CuO was observed below 700 °C [55].
Based on the above analysis, Cu(OH)2, CuO, CuCO3, and
Cu2(OH)2CO3 were included in the precipitates.

4. Conclusions

The influences of coexisting inorganic ions on EDTA-Cu decom-


plexation and Cu2+ release from complexes were evaluated in a non-
thermal plasma oxidation coupled with chemical precipitation system.
The presence of Cl− promoted EDTA-Cu decomplexation due to the
complexing competitive effect of Cl− with copper and compensation
Fig. 15. FTIR, Raman and TG-DTG analysis of the precipitates after 50 min effect of the formation of Cl-containing oxidizing species. CO32− sig-
treatment (a. FTIR; b. Raman; c. TG-DTG). nificantly inhibited EDTA-Cu decomplexation because of the loss of %
OH. The concomitant NO3− promoted EDTA-Cu decomplexation due to
the compensation effect of the formation of O and NO3% radical. The
employed to characterize the precipitates. Fig. 13(a)–(d) depicts the
replacement effect and Fenton-like effect of Fe3+ significantly pro-
EDX mapping of the precipitates, in which C, O, and Cu were observed.
moted EDTA-Cu decomplexation, while only the replacement effect was
This finding demonstrated that the precipitates mainly included C, O,
observed in the presence of Ni2+. These influences of coexisting in-
and Cu elements, and thus copper oxides and Cu-based carbonates
organic ions were also analyzed via TOC and Cu2+ release. 1O2, %OH,
might exist in the precipitates. A typically SEM photograph of the
and %O2− played crucial roles in EDTA-Cu decomplexation, and O3 and
precipitates is shown in Fig. 14, where precipitates with different
%O also participated in the process. Ethanamine, acetamide, glycolic
shapes were observed; this result further indicated that different Cu-
acid, formamide, ethylene glycol, CH3COOH, and butanedioic acid
based compounds simultaneously existed in the precipitates.
were observed after EDTA-Cu decomplexation. Attacks of reactive
Fig. 15(a) depicts the FTIR spectrum of the precipitates. The peaks
substances on CeN bonds in EDTA-Cu triggered its decomplexation
at 1436, 1338, and 1130 cm−1 were associated with Cu(II)-based car-
process, leading to the generation of small molecular fragments and
bonates, eC]O symmetric vibration, and CeO stretching vibration,
Cu2+ release. Cu2CO3(OH)2, CuO, Cu(OH)2 and CuCO3 were observed
respectively; the two absorption bands at 877 and 781 cm−1 were as-
in the precipitates based on chemical composition analysis.
sociated with the bidentate carbonates; and the peaks at approximately
617 and 516 cm−1 corresponded to Cu(II)eO vibration [47–49]. These
Acknowledgments
results further confirmed that copper oxides and Cu-based carbonates
were formed in the precipitates. In addition, the eOH group of bending
The National Natural Science Foundation of China (51608448,

495
Y. Cao et al. Chemical Engineering Journal 362 (2019) 487–496

21737003), Young Talent Cultivation Scheme Funding of Northwest A& oxidation of methyl violet with gliding arc plasma discharge, Plasma Chem. Plasma
F University (Z109021802), and Science and Technology Innovation Process. 33 (2013) 737–749.
[28] D.R. Merouani, F. Abdelmalek, F. Taleb, M. Martel, A. Semmoud, A. Addou, Plasma
Project of Yangling (2018SF-03) supported this research. treatment by gliding arc discharge of dyes/dye mixtures in the presence of in-
organic salts, Arab. J. Chem. 8 (2015) 155–163.
References [29] Y.Q. Zhang, J.F. Zhang, Y.J. Xiao, V.W.C. Chang, T.T. Lim, Kinetic and mechanistic
investigation of azathioprine degradation in water by UV, UV/H2O2 and UV/per-
sulfate, Chem. Eng. J. 302 (2016) 526–534.
[1] H. Sawai, I.M.M. Rahman, M. Fujita, N. Jii, T. Wakabayashi, Z.A. Begum, T. Maki, [30] Z. Xu, C. Shan, B.H. Xie, Y. Liu, B.C. Pan, Decomplexation of Cu(II)-EDTA by UV/
S. Mizutani, H. Hasegawa, Decontamination of metal-contaminated waste foundry persulfate and UV/H2O2: efficiency and mechanism, Appl. Catal. B-Environ. 200
sands using an EDTA-NaOH-NH3 washing solution, Chem. Eng. J. 296 (2016) (2017) 439–447.
199–208. [31] F.M. Huang, L. Chen, H.L. Wang, Z.C. Yan, Analysis of the degradation mechanism
[2] P. Haapea, T. Tuhkanen, Integrated treatment of PAH contaminated soil by soil of methylene blue by atmospheric pressure dielectric barrier discharge plasma,
washing, ozonation and biological treatment, J. Hazard. Mater. 136 (2006) Chem. Eng. J. 162 (2010) 250–256.
244–250. [32] Y. Katsumura, P.Y. Jiang, R. Nagaishi, T. Oishi, K. Ishigure, Pulse radiolysis study of
[3] R.D. Villa, A.G. Trovo, R.F.P. Nogueira, Soil remediation using a coupled process: aqueous nitric acid solutions. Formation mechanism, yield, and reactivity of NO3
soil washing with surfactant followed by photo-Fenton oxidation, J. Hazard. Mater. radical, J. Phys. Chem. 95 (1991) 4435–4439.
174 (2010) 770–775. [33] Y.J. Xiao, L.F. Zhang, J.Q. Yue, R.D. Webster, T.T. Lim, Kinetic modeling and energy
[4] X.F. Guo, G.X. Zhang, Z.B. Wei, L.P. Zhang, Q.S. He, Q.T. Wu, T.W. Qian, Mixed efficiency of UV/H2O2 treatment of iodinated trihalomethanes, Water Res. 75
chelators of EDTA, GLDA, and citric acid as washing agent effectively remove Cd, (2015) 259–269.
Zn, Pb, and Cu from soils, J. Soils Sediments 18 (2018) 835–844. [34] M.I. Litter, Heterogeneous photocatalysis transition metal ions in photocatalytic
[5] P. Van der Maas, S. Peng, B. Klapwijk, P. Lens, Enzymatic versus nonenzymatic systems, Appl. Catal. B-Environ. 23 (1999) 89–114.
conversions during the reduction of EDTA-chelated Fe(III) in BioDeN(x) reactors, [35] B. Nowack, J. Lutzenkirchen, P. Behra, L. Sigg, Modeling the adsorption of metal-
Environ. Sci. Technol. 39 (2005) 2616–2623. EDTA complexes onto oxides, Environ. Sci. Technol. 30 (1996) 2397–2405.
[6] D. Fabbri, A. Crime, M. Davezza, C. Medana, C. Baiocchi, A.B. Prevot, E. Pramauro, [36] S. Aguilar, D. Rosado, J. Moreno-Andres, L. Cartuche, D. Cruz, A. Acevedo-Merino,
Surfactant-assisted removal of swep residues from soil and photocatalytic treatment E. Nebot, Inactivation of a wild isolated Klebsiella pneumoniae by photo-chemical
of the washing wastes, Appl. Catal. B-Environ. 92 (2009) 318–325. processes: UV-C, UV-C/H2O2 and UV-C/H2O2/Fe3+, Catal. Today 313 (2018)
[7] L.K. Malinen, R. Koivula, R. Harjula, Removal of radiocobalt from EDTA-complexes 94–99.
using oxidation and selective ion exchange, Water Sci. Technol. 60 (2009) [37] A.A. Joshi, B.R. Locke, P. Arce, W.C. Finney, Formation of hydroxyl radicals, hy-
1097–1101. drogen-peroxide and aqueous electrons by pulsed streamer corona discharge in
[8] X. Zhao, L.B. Guo, B.F. Zhang, H.J. Liu, J.H. Qu, Photoelectrocatalytic oxidation of aqueous-solution, J. Hazard. Mater. 41 (1995) 3–30.
Cu-II-EDTA at the TiO2 electrode and simultaneous recovery of Cu-II by electro- [38] P.S. Rao, E. Hayon, Redox potentials of free radicals. IV. Superoxide and hydro-
deposition, Environ. Sci. Technol. 47 (2013) 4480–4488. peroxy radicals. O2· and ·HO2, J. Phys. Chem. 79 (1975) 397–402.
[9] S.Y. Lan, Y. Xiong, S.H. Tian, J.X. Feng, T.Y. Xie, Enhanced self-catalytic de- [39] H. Görner, Photoprocesses of p-naphthoquinones and vitamin K1: effects of alcohols
gradation of CuEDTA in the presence of H2O2/UV: evidence and importance of Cu- and amines on the reactivity in solution, Photochem. Photobiol. Sci. 3 (2004)
peroxide as a photo-active intermediate, Appl. Catal. B-Environ. 183 (2016) 71–78.
371–376. [40] D.D. Dionysiou, Efficient removal of microcystin-LR by UV-C/H2O2 in synthetic and
[10] T.C. Wang, G.Z. Qu, J.Y. Ren, Q.H. Yan, Q.H. Sun, D.L. Liang, S.B. Hu, Evaluation of natural water samples, Water Res. 46 (2012) 1501–1510.
the potentials of humic acid removal in water by gas phase surface discharge [41] X. Zhao, L.B. Guo, C.Z. Hu, H.J. Liu, J.H. Qu, Simultaneous destruction of Nickel
plasma, Water Res. 89 (2016) 28–38. (II)-EDTA with TiO2/Ti film anode and electrodeposition of nickel ions on the
[11] Y. Cao, G.Z. Qu, T.F. Li, N. Jiang, T.C. Wang, Review on reactive species in water cathode, Appl. Catal. B-Environ. 144 (2014) 478–485.
treatment using electrical discharge plasma: formation, measurement, mechanisms [42] X.F. Huang, Y. Xu, C. Shan, X.C. Li, W.M. Zhang, B.C. Pan, Coupled Cu(II)-EDTA
and mass transfer, Plasma Sci. Technol. 20 (2018) 103001–103017. degradation and Cu(II) removal from acidic wastewater by ozonation: performance,
[12] N. Jiang, L.J. Guo, C. Qiu, Y. Zhang, K.F. Shang, N. Lu, J. Li, Y. Wu, Reactive species products and pathways, Chem. Eng. J. 299 (2016) 23–29.
distribution characteristics and toluene destruction in the three-electrode DBD re- [43] C. Oviedo, D. Contreras, J. Freer, J. Rodriguez, Fe(III)-EDTA complex abatement
actor energized by different pulsed modes, Chem. Eng. J. 350 (2018) 12–19. using a catechol driven Fenton reaction combined with a biological treatment,
[13] T.C. Wang, H.Z. Jia, X.T. Guo, T.J. Xia, G.Z. Qu, Q.H. Sun, X.Q. Yin, Evaluation of Environ. Technol. 25 (2004) 801–807.
the potential of dimethyl phthalate degradation in aqueous using sodium percar- [44] F. Ju, Y.Y. Hu, Removal of EDTA-chelated copper from aqueous solution by interior
bonate activated by discharge plasma, Chem. Eng. J. 346 (2018) 65–76. microelectrolysis, Sep. Purif. Technol. 78 (2011) 33–41.
[14] S.F. Tang, D.L. Yuan, Y.D. Rao, N. Li, J.B. Qi, T.Z. Cheng, Z.T. Sun, J.M. Gu, [45] S.S. Lee, H.W. Bai, Z.Y. Liu, D.D. Sun, Green approach for photocatalytic Cu(II)-
H.M. Huang, Persulfate activation in gas phase surface discharge plasma for sy- EDTA degradation over TiO2: toward environmental sustainability, Environ. Sci.
nergetic removal of antibiotic in water, Chem. Eng. J. 337 (2018) 446–454. Technol. 49 (2015) 2541–2548.
[15] A.M. Lietz, M.J. Kushner, Air plasma treatment of liquid covered tissue: long [46] F. Liu, C. Shan, X.L. Zhang, Y.Y. Zhang, W.M. Zhang, B.C. Pan, Enhanced removal of
timescale chemistry, J. Phys. D-Appl. Phys. 49 (2016) 425204–425225. EDTA-chelated Cu(II) by polymeric anion-exchanger supported nanoscale zero-va-
[16] P.J. Bruggeman, M.J. Kushner, B.R. Locke, et al., Plasma-liquid interactions: a re- lent iron, J. Hazard. Mater. 321 (2017) 290–298.
view and roadmap, Plasma Sources Sci. Technol. 25 (2016) 053002–053060. [47] A. Bernson, J. Lindgren, W.W. Huang, R. Frech, Coordination and conformation in
[17] E.C. Neyts, M. Yusupov, C.C. Verlackt, A. Bogaerts, Computer simulations of Peo, Pegm and Peg systems containing lithium or lanthanum triflate, Polymer 36
plasma-biomolecule and plasma-tissue interactions for a better insight in plasma (1995) 4471–4478.
medicine, J. Phys. D-Appl. Phys. 47 (2014) 293001–293017. [48] Y. Qiao, Y.Y. Lin, S. Liu, S.F. Zhang, H.F. Chen, Y.J. Wang, Y. Yan, X.F. Guo,
[18] D.B. Graves, Mechanisms of plasma medicine: coupling plasma physics, biochem- J.B. Huang, Metal-driven hierarchical self-assembled zigzag nanoarchitectures with
istry and biology, IEEE Trans. Radiat. Plasma Med. Sci. 1 (2017) 281–292. electrical conductivity, Chem. Commun. 49 (2013) 704–706.
[19] Y.J. Liu, Simultaneous oxidation of phenol and reduction of Cr (VI) induced by [49] I. Schrader, L. Wittig, K. Richter, H. Vieker, A. Beyer, A. Golzhauser, A. Hartwig,
contact glow discharge electrolysis, J. Hazard. Mater. 168 (2009) 992–996. P. Swiderek, Formation and structure of copper(II) oxalate layers on carboxy-ter-
[20] Z.G. Ke, Q. Huang, H. Zhang, Z.L. Yu, Reduction and removal of aqueous Cr(VI) by minated self-Assembled monolayers, Langmuir 30 (2014) 11945–11954.
glow discharge plasma at the gas-solution interface, Environ. Sci. Technol. 45 [50] B. Peng, T.T. Song, T. Wang, L.Y. Chai, W.C. Yang, X.R. Li, C.F. Li, H.Y. Wang, Facile
(2011) 7841–7847. synthesis of Fe3O4@Cu(OH)(2) composites and their arsenic adsorption application,
[21] T.C. Wang, Y. Cao, G.Z. Qu, Q.H. Sun, T.J. Xia, X.T. Guo, H.Z. Jia, L.Y. Zhu, Novel Chem. Eng. J. 299 (2016) 15–22.
Cu(II)-EDTA decomplexation by discharge plasma oxidation and coupled Cu re- [51] Y.L. Deng, A.D. Handoko, Y.H. Du, S.B. Xi, B.S. Yeo, In situ Raman spectroscopy of
moval by alkaline precipitation: underneath mechanisms, Environ. Sci. Technol. 52 copper and copper oxide surfaces during electrochemical oxygen evolution reac-
(2018) 7884–7891. tion: identification of Cu-III oxides as catalytically active species, ACS Catal. 6
[22] M. Muthukumar, N. Selvakumar, Studies on the effect of inorganic salts on deco- (2016) 2473–2481.
louration of acid dye effluents by ozonation, Dyes Pigm. 62 (2004) 221–228. [52] R.L. Frost, A Raman spectroscopic study of selected minerals of the rosasite group,
[23] A.C. Silva, J.S. Pic, G.L. Sant'Anna, M. Dezotti, Ozonation of azo dyes (Orange II and J. Raman Sectors 37 (2006) 910–921.
Acid Red 27) in saline media, J. Hazard. Mater. 169 (2009) 965–971. [53] D.D. La, H.P.N. Thi, Y.S. Kim, A. Rananaware, S.V. Bhosale, Facile fabrication of Cu
[24] A.R. Tehrani-Bagha, N.M. Mahmoodi, F.M. Menger, Degradation of a persistent (II)-porphyrin MOF thin films from tetrakis(4-carboxyphenyl) porphyrin and Cu
organic dye from colored textile wastewater by ozonation, Desalination 260 (2010) (OH)(2) nanoneedle array, Appl. Surf. Sci. 424 (2017) 145–150.
34–38. [54] R.L. Frost, A.J. Locke, M.C. Hales, W.N. Martens, Thermal stability of synthetic
[25] X.Y. Yu, J.R. Barker, Hydrogen peroxide photolysis in acidic aqueous solutions aurichalcite implications for making mixed metal oxides for use as catalysts, J.
containing chloride ions. I. chemical mechanism, J. Phys. Chem. A 107 (2003) Therm. Anal. Calorim. 94 (2008) 203–208.
1313–1324. [55] C.H. Tsai, P.H. Fei, C.M. Lin, S.L. Shiu, CuO and CuO/Graphene nanostructured thin
[26] S.N. Ramjaun, R.X. Yuan, Z.H. Wang, J.S. Liu, Degradation of reactive dyes by films as counter electrodes for Pt-free dye-sensitized solar cells, Coatings 8 (2018)
contact glow discharge electrolysis in the presence of Cl- ions: kinetics and AOX 21–33.
formation, Electrochim. Acta 58 (2011) 364–371.
[27] Y.N. Liu, S.F. Zhu, H. Tian, M. Zhou, J. Miao, Effect of inorganic ions on the

496

You might also like