You are on page 1of 9

Chemosphere 329 (2023) 138651

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Degradation of trichloroethylene by double dielectric barrier discharge


(DDBD) plasma technology: Performance, product analysis and acute
biotoxicity assessment
Xu-Rui Hu a, b, Yong-Chao Wang a, b, Zhen Tong a, b, Can Wang a, b, *, Er-Hong Duan c,
Meng-Fei Han c, Hsing-Cheng Hsi d, Ji-Guang Deng e
a
School of Environmental Science and Engineering, Tianjin University, Tianjin, 300072, China
b
Tianjin Key Lab of Indoor Air Environmental Quality Control, Tianjin, 300072, China
c
School of Environmental Science and Engineering, Hebei University of Science and Technology, Shijiazhuang, Hebei, 050018, China
d
Graduate Institute of Environmental Engineering, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd, Taipei, 106, Taipei, Taiwan
e
College of Environmental and Energy Engineering, Beijing University of Technology, Beijing, 100124, China

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

• Effects of some operational conditions


on TCE removal by DDBD were
evaluated.
• Energy yield and kinetics of TCE degra­
dation by DDBD were calculated.
• The possible decomposition pathways of
DDBD in this system were proposed.
• Residual toxicity after DDBD decompo­
sition was significantly increased.

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

Handling Editor: E. Brillas Trichloroethylene is carcinogenic and poorly degraded by microorganisms in the environment. Advanced
Oxidation Technology is considered to be an effective treatment technology for TCE degradation. In this study, a
Keywords: double dielectric barrier discharge (DDBD) reactor was established to decompose TCE. The influence of different
Non-thermal plasma condition parameters on DDBD treatment of TCE was investigated to determine the appropriate working con­
Double dielectric barrier discharge
ditions. The chemical composition and biotoxicity of TCE degradation products were also investigated. Results
Trichloroethylene
showed that when SIE was 300 J L− 1, the removal efficiency could reach more than 90%. The energy yield could
Energy yield
Biotoxicity reach 72.99 g kWh− 1 at low SIE and gradually decreased with the increase of SIE. The k of the Non-thermal
plasma (NTP) treatment of TCE was about 0.01 L J− 1. DDBD degradation products were mainly poly­
chlorinated organic compounds and produced more than 373 mg m− 3 ozone. Moreover, a plausible TCE
degradation mechanism in the DDBD reactors was proposed. Lastly, the ecological safety and biotoxicity were
evaluated, indicating that the generation of chlorinated organic products was the main cause of elevated acute
biotoxicity.

* Corresponding author. School of Environmental Science and Engineering, Tianjin University, Tianjin, 300072, China.
E-mail address: wangcan@tju.edu.cn (C. Wang).

https://doi.org/10.1016/j.chemosphere.2023.138651
Received 23 January 2023; Received in revised form 7 April 2023; Accepted 7 April 2023
Available online 12 April 2023
0045-6535/© 2023 Elsevier Ltd. All rights reserved.
X.-R. Hu et al. Chemosphere 329 (2023) 138651

(Santos et al., 2020). Nevertheless, the biotoxicity of the plasma-treated


Nomenclature products of these VOCs has rarely been reported.
In this study, TCE decomposition and product generation were
NTP non-thermal plasma evaluated in a DDBD reactor. Ozone generation under different condi­
DDBD double dielectric barrier discharge tions was also evaluated. The effects of SIE (specific input energy), inlet
RE removal efficiency [%] TCE concentration, and inlet gas flow rate on the removal efficiency,
SIE specific input energy [J L− 1] removal rate, and energy yield by DDBD were investigated. In addition,
EY energy yield [g kWh− 1] the kinetics and the reaction pathway were examined. Finally, the bio­
Pw discharge power [W] toxicity of the by-product(s) was also evaluated for the possibility of
Cin inlet concentration [ppm] subsequent biodegradation.
Cout outlet concentration [ppm]
DOM dissolved organic matter 2. Material and methods
COx CO2 and CO
RH relative humidity 2.1. Experimental set-up

The schematic diagram of the experimental set-up is shown in Fig. 1.


The experimental apparatus is composed of a TCE generator, a DDBD
1. Introduction unit, and analytical equipment. Simulated waste gas was obtained by
injecting a certain amount of TCE into the dry airflow. One bubbler
Currently, trichloroethylene (TCE) is widely used in coatings, contained demineralized water for defined humidification of the
cleaning technology, daily chemical production, and other industries as airflow, another bubbler was used to mix the two strands of airflow and
a good solvent. In the past few decades, people have leaked TCE into the maintain a stable airflow. The DDBD unit was composed of a DDBD
environment during production and use, which have made it widely reactor and a high-voltage AC power supply. The gas products were
distributed in the atmosphere, soil, and water environments (Deng et al., analyzed by Gas chromatography and other methods.
2022). TCE is highly volatile, and persistent in the environment due to The DDBD reactor used quartz tubes as dielectric with two different
its non-biodegradability (Lei et al., 2018). It has been proven to be one of sizes. The inner diameters were 6 mm and 13 mm, and the thicknesses
the most hazardous volatile organic compounds (VOCs), being both were 1 mm and 1.5 mm, respectively. A stainless steel bar (6 mm in
carcinogenic and mutagenic. At the same time, TCE can cause diameter) was employed as the high-voltage electrode, and a stainless
dysfunction of the human nervous system and cause vertigo, nausea, steel mesh wrapping outside the outer quartz glass tube served as the
sleepiness, fatigue, cardiac failure, etc (USEPA and, 2007). Therefore, it ground electrode. The discharge length was 8 cm. The drive power
is urgent to develop technologies that are feasible to control this supply for the reactor is a high-voltage AC power supply (CTP-2000 K,
pollutant. China). The output voltage is 0–30 kV, and the center frequency is about
In recent years, various technologies have been widely used to con­ 10 kHz. The current and voltage were measured and displayed by a
trol TCE (Zhang et al., 2016). However, conventional techniques are not digital storage oscilloscope (TDS2024C, America). Plasma discharge
energy efficient for the removal of this pollutant (Tang et al., 2013). power was calculated by the Lissajous pattern method. The detailed in­
Biological methods have been considered as a cost-effective technology formation was given in Supporting Materials (S1). The experiment was
because of their convenient maintenance, low operating costs, and low conducted under room temperature and atmospheric pressure
secondary pollution (Wang et al., 2023). But its application is limited by conditions.
the low solubility and low biodegradability of TCE. For the past few
years, advanced oxidation processes (AOPs) have been documented as
effective pretreatment technologies (Wang et al., 2009; Paz, 2010; Li 2.2. Experimental procedure
et al., 2018). Alcantara-Garduno et al. (2005) have developed a com­
bination process of ozone and acetic acid flushing to decompose TCE. Experiments were performed by varying the inlet gas flow rate to
Shen and Ku (2002) have evaluated direct photolysis, the UV/TiO2, and obtain different initial concentrations and residence times. The perfor­
UV/O3 processes to decompose TCE, and the result showed that the mance was evaluated by the formulas below:
UV/TiO2 process achieved the best removal. Li et al. (2019) have
Cin − Cout
fabricated a bentonite supported Fe/Ni bimetallic nanoparticle (BNF) Removal efficiency (%) = × 100 (1)
Cin
for peroxymonosulfate (PMS) activation to degrade TCE in ground­
water, and complete degradation of 0.1 mM TCE was achieved in 25
min. However, the long residence time required by these technologies
limits their practical application.
In non-thermal plasma (NTP) technology, the input electrical energy
is transferred to generate energetic high-energy electrons, and highly
reactive species such as ozone, atomic oxygen (⋅O), and hydroxyl radical
(⋅OH) to decompose the pollutants (Schiavon et al., 2017). Thus, the
pollutant can be converted into CO2, H2O, and other products. Jiang
et al. (2015) have evaluated the production and biodegradability of
mixed VOCs by DBD treatment. The results showed that styrene and
o-xylene were converted into other BTEX and that their biodegradability
was improved. Guo et al. (2021) have established a coaxial double
dielectric barrier discharge reactor to degrade low concentrations of
ethyl acetate. The results showed that when the inlet concentration was
low, the degradation efficiency could reach 96.4%, and the degradation
products included alcohols, aldehydes, and acids. However, the forma­
tion of unwanted by-products such as NOX, O3, other VOCs, organic
aerosol, etc., may increase the overall toxicity of the treated pollutant Fig. 1. Experimental platform for DDBD treatment of TCE.

2
X.-R. Hu et al. Chemosphere 329 (2023) 138651

( ) Q × (Cin − Cout ) 3. Results and discussion


Removal rate g m− 3 h− 1 = (2)
V
( ) 3.1. Effect of operation conditions on TCE removal
Pw
SIE J L− 1 = (3)
Q The effects of inlet TCE concentration and SIE on RE and removal
rate were shown in Fig. 2a and b, respectively. The RE of TCE was
( ) 3.6 × MTCE × (Cin − Cout )
EY g kWh− 1 = (4) enhanced by the increase in the SIE. The RE decreased from 90% to 77%
22.4 × SIE with the TCE concentration increasing from 151 ppm to 933 ppm when
SIE was 203 J L− 1. This is since the amount of active species produced is
where Cin and Cout are concentrations of TCE before and after plasma
determined by the discharge power. Given a certain discharge power,
treatment; Q is the gas flow rate; V is the effective volume of the DDBD
the number of reactive species generated by DDBD discharge is constant.
reactor; Pw is discharge power (W); EY is energy yield; MTCE is the
The increase in TCE concentration results in fewer reactive species and
relative molecular mass of TCE; and 22.4 is the molar volume of gas. The
the sharing of high-energy electrons (Vandenbroucke et al., 2011). As
high-voltage AC power supply can directly obtain the input voltage and
shown in Fig. 2b, the removal rate of TCE increased from about 8 × 103
current of the reactor and then obtain the discharge power of the
g m− 3 h− 1 to 3.6 × 104 g m− 3 h− 1 with the increase of SIE. Jiang et al.
reactor. Duplicate experiments were performed on samples of the same
(2015) also found that the removal amount of VOCs increased with an
conditions during the experiment three times to ensure sampling
increase in inlet gas concentration.
accuracy.
Fig. 2c and d have shown the effects of inlet gas flow rates and SIE on
RE and removal rate for a TCE inlet concentration of 450 ± 25 ppm. The
2.3. Chemical identification of TCE and degradation products
RE of TCE increased with the increase in the SIE and decreased with the
increase in the inlet flow rate when the SIE was constant. However, the
A gas chromatograph (GC7900, TianMei, China) was used to analyze
removal rate of TCE increased with the increase of the inlet flow rate.
TCE concentrations at the inlet and outlet of the reactor. The RH of the
When the TCE inlet flow rate increased from 0.6 L min− 1 to 1.6 L min− 1,
inlet gas was measured by a dust particle counter equipped with a hu­
the RE decreased from 93% to 69% when SIE was 170 J L− 1. The phe­
midity sensor (DT-9880, CEM, China). The iodometry method was used
nomenon is attributed to the increased number of pollutant molecules
to measure the concentration of ozone (Dou et al., 2013). A Gas
passing through the reactor per unit of time (Chavadej et al., 2007).
Chromatography-Mass Spectrometry (GC-MS7980A-5975C, Agilent)
Meanwhile, the number of pollutant molecules per unit volume in­
was used to analyze degradation products. Exhaust gas is successively
creases, causing the probability of collision between pollutant molecules
passed into two absorption bottles filled with deionized water. The
and active species to increase, which increases the removal amount of
organic products and ionic compounds were detected by the Total
the pollutant. As shown in Fig. 2d, the removal rate of TCE increased
Organic Carbon Analyzer (TOC-L, Shimadzu, Japan) and Ion Chroma­
from 1.4 × 104 g m− 3 h− 1 to 3.1 × 104 g m− 3 h− 1 when the inlet flow rate
tography (ICS-1100, THERMO, America), respectively.
increased from 0.6 L min− 1 to 1.6 L min− 1 when SIE was 181 J L− 1.
According to Eqs. (1) and (2), we can conclude that the removal rate
2.4. Biotoxicity assessment of degradation products
is determined by the inlet gas concentration, RE, inlet gas flow rate, and
reaction volume. Since RE increases with SIE, for a certain inlet con­
TCE is a highly toxic and refractory organic pollutant. The toxicity of
centration, inlet flow rate, and reaction volume, the removal rate of TCE
the intermediate product of pollutant degradation may become higher
increases with SIE. Meanwhile, the removal rate is proportional to inlet
than the parent compound (Wang et al., 2008). The acute biotoxicity of
concentration and inlet flow and inversely proportional to reaction
the products was measured by a luminescent bacteria method. This
volume. Therefore, when SIE is constant, the removal rate increases with
method is widely used to evaluate the acute toxicity of water. The
the increase in inlet concentration and inlet gas flow rate.
principle of this method is that the luminescence of luminescent bacteria
is its normal metabolic activity, and the luminescence intensity is con­
stant under certain conditions. When the luminescent bacteria came into 3.2. Kinetics of TCE degradation by DDBD
contact with foreign substances (inorganic, organic toxicants, bacterio­
static, bactericidal, etc.), the luminescence intensity changed. In a Knowing the reaction rate constant is vital for evaluating the ease of
certain range, the luminescence intensity change was related to the VOC degradation and predicting the elimination of VOCs in NTP re­
concentration of the subject and also related to the toxicity of the sub­ actors. The process of converting DDBD to TCE takes place in a cylin­
stance. In order to visually compare the acute toxicity of the samples and drical quartz glass reactor, and the gas flow can be assumed to be a plug
increase the comparability of the results of different batches of tests, the flow (Karatum and Deshusses, 2016). The elementary physical and
acute toxicity of the samples in this study was characterized by the chemical processes in the plasma system are complex, and the radicals
concentration of the corresponding reference poison Zn2+. The exposure and the intermediates in the process are hard to measure directly.
time of the experiment was set at 15 min. The luminescence inhibition In this study, a simplified model (Eq. (5), (6), and (7)) was used to
rate at this time was equivalent to the zinc concentration under this illustrate the degradation process of TCE by DDBD. These reactions are
inhibition rate, so as to compare the influence of the product on the all related by the number of high-energy electrons and active species,
luminescence inhibition rate (Wang et al., 2008). At the same time, which is correlated with specific input energy (SIE). So all the reaction
different substances (sodium hydroxide and sodium thiosulfate, etc.) process can be reduced to a simple equation. The model is generally
were added to analyze the effects of different properties of products applicable when the VOC removal does not exceed 95%. The slope of the
(acidity and oxidation) on luminescence inhibition. straight line (1/β) is (minus) the reaction rate constant k (L J− 1).
At the same time, the Toxicity Estimation Software Tool (T.E.S.T)
was used to predict the biotoxicity of degradation products or possible (5)
k
High − energy electrons / active species + [VOC]→products
degradation products. T.E.S.T mainly obtains the toxicity of substances
by analyzing the known QSAR (Quantitative Structure-Activity Rela­ [VOC]
= e− SIE/β
(6)
tionship) model, the special literature published by the US EPA and ITC [VOC]0
(Interagency Testing Committee). ( )
[VOC] 1
ln = SIE × − (7)
[VOC]0 β

3
X.-R. Hu et al. Chemosphere 329 (2023) 138651

Fig. 2. (a. Effect of the initial inlet concentration on TCE RE; b. Effect of the initial inlet concentration on TCE removal rate; c. Effect of the inlet gas flow rates on TCE
RE; d. Effect of the inlet gas flow rates on TCE removal rate.)

Fig. 3. Kinetics analysis and the energy yield of TCE conversion in the DDBD system.

4
X.-R. Hu et al. Chemosphere 329 (2023) 138651

A linear relationship between the data of ln[VOC]/[VOC]0 and SIE is From Section 3.2, we can concluded that β is a constant value (1/k),
depicted in Fig. 3, indicating that the conversion of TCE by DDBD can be when the TCE inlet gas concentration is certain, the equation can be
described by primary reaction kinetics. When the degradation rate simplified as:
constants of TCE were compared under different reaction conditions, the
1 − e(− SIE/β)
results showed that they ranged from 0.0097 L J− 1 to 0.01319 L J− 1 in EY = m × (9)
Table S1, which was in the same order of magnitude as the degradation SIE
rate constants of TCE obtained by Han and Oda (2007). However, these Since m is a positive number greater than 1, according to Eq. (9), it
are different from the degradation rate constants of other VOCs, such as can be inferred that EY decreases as SIE increases.
0.0076, 0.0044, 0.0033, 0.0022, and 0.0018 L J− 1 for n-hexane, MTBE, Li et al. (2020) compared the EY of toluene treated with SDBD and
toluene, benzene, and MEK, respectively, indicating that TCE is more DDBD and found that the EY of toluene dropped from 2.49 to 1.53 g
suitable for treatment with NTP technology (Karatum and Deshusses, kWh− 1 in SDBD. Nguyen et al. (2020) evaluated the EY of toluene and
2016). This may be due to the low bond energy of C–Cl, which leads to MEK in the surface dielectric barrier discharge reactors. EY increased as
the easy conversion of TCE. However, TCE is not easy to be completely the inlet concentration increased from 20 to 100 ppm, while decreasing
mineralized due to the high bond energy of C=C but forms other as the SIE increased. Overall, DDBD is economically suitable for pro­
chlorine-containing organics (Jiang et al., 2022). cessing high-concentration and high-gas-flow TCE gas under low SIE
(a. Effect of inlet concentration on kinetics; b. Effect of inlet gas flow conditions. Table S2 compared the EY of the single NTP technology for
rate on kinetics; c. Effect of inlet concentration on energy yield; d. Effect the treatment of different VOCs. As shown in Table S2, the EY of TCE is
of inlet gas flow rate on energy yield.) the highest among these VOCs and is an order of magnitude higher than,
for example, toluene and chlorobenzene. Despite the differences in the
size of the plasma reactor and the form of discharge, it can also be
3.3. EY of TCE conversion by DDBD explained to some extent that TCE is suitable for treatment using NTP
technology.
Energy yield (EY) is one of the most important criteria for evaluating
a process and is used to measure the economics of an industrial plant.
Fig. 3c–d shows the effects of different process conditions on EY. The 3.4. Identification of inorganic (by)products
results showed that EY ranged from 4.91 to 72.99 g kWh− 1 and
decreased as the SIE increased. The EY was gradually flattened by the The anions absorbed by the deionized water were analyzed by Ion
SIE as the inlet gas flow increased. This is consistent with the results of Chromatography. And the pH of the absorption solution was also tested.
other studies. The following equation can be deduced from Eqs. (4) and The results are shown in Fig. 4a–b.
(6): The experimental results showed that with the increase of the SIE,
(
3.6 × MTCE × Cin × 1 − e(− SIE/β)
) the concentration of chloride and nitrate in the absorbed solution
EY = (8) gradually increased after the treatment with TCE, the concentration of
22.4 × SIE
nitrite was very low and remained unchanged, and the pH of the

Fig. 4. Identification of products and distribution. (a. Ion concentration and pH of water samples after TCE treatment; b. Ion concentration and pH of water samples
after the plasma process without the addition of TCE; c. Mass distribution of the carbon element; d. Mass distribution of the chlorine element.)

5
X.-R. Hu et al. Chemosphere 329 (2023) 138651

absorbed solution also gradually decreased. This is because TCE was on acidic pH paper, the color of the paper first turned red and then
decomposed and dechlorinated after the DDBD treatment to produce white, presumably as chlorine water formed by dissolving chlorine gas
HCl. With the increase of SIE, the RE of TCE gradually increased, and the in condensate vapor. Han and Oda (2007) and Vandenbroucke et al.
generation of Cl− also increased gradually. Due to the production of (2011) demonstrated that the production of TCE decomposition by
nitrogen oxides (NOx), NO−3 was produced in the water sample (Shahna plasma treatment contained chlorine gas.
et al., 2017). The undesirable by-product NOx produced by the plasma
discharge process dissolves in water and reacts with it to form NO−3 and 3.6. Proposed reaction pathways for TCE degradation by DDBD
NO−2 , while NO−2 is extremely unstable under acidic conditions and will
react to form NO−3 . Also, the ozone generated by the DDBD oxidized the Previous studies (Han and Oda., 2007; Vandenbroucke et al., 2011)
NO−2 generated in the water to produce NO−3 . The NOx produced was have investigated the plasma decomposition of TCE in the gas phase and
mainly NO2, and the result is shown in Fig. S3. The concentration of NO2 demonstrated that dichloroacetyl chloride (DCAC), tri­
increased with the increase of SIE from 167 mg m− 3 to 244 mg m− 3. The chloroacetaldehyde (TCAA), and phosgene were presented in the (by)
results were consistent with Han et al.’s (2020) finding that the NOx product. DCAC was the main product in both single plasma and plasma
produced by the dielectric discharge plasma was dominated by NO2, catalytic assisted systems. In this study, the GC-MS of the substances
while the NO2 concentration produced in this experiment was much were analyzed and the result were shown in Fig. S5. Phosgene and other
lower than that. This may be caused by the different parameters, such as polychlorinated organic compounds was detected in the absorption so­
reactor size and input power supply. lution. Electron-impact dissociation of H2O and O2 can generate large
amounts of hydroxyl radicals and atomic oxygen species, respectively, in
3.5. Identification of organic (by)products humid air. From Eqs. 10–14 (NIST, 2011), we can observe that the k of
the reaction between TCE and high-energy electrons is the highest.
Fig. 4c shows the results of the DDBD reactor’s carbon element mass Meanwhile, the dissociation energy of the C=C bond (6.3 eV) and C–H
distribution. The mass of the carbon absorbed by the water sample bond (4.3 eV) are higher than that of C–Cl bond (3.5 eV). So the first
(Fig. S4) was increased and then decreased with increasing SIE (corre­ reaction of TCE should be to react with high-energy electrons to
sponding to “DOM” in Fig. 4c). And the COX content increased from 27% generate ⋅Cl. Then, ⋅Cl or ROS/RNS continue to undergo subsequent
to 65%. Because as the SIE increases, the number of active species reactions with TCE to generate chlorinated hydrocarbons, oxygenated
produced increases, allowing for a more complete decomposition of TCE chlorinated organics, and nitrogen-containing chlorinated organics. The
into COX (Jiang et al., 2013). Jiang et al. (2015) used a DBD reactor to reaction between radical Cl and TCE could generate C2Cl3, HCl, and
treat a mixture of VOCs. On increasing SIE, CO and CO2 reached 27.2% C2HCl4. DCAC would be formed by the reaction of Cl radicals and the
and 40%, respectively on increasing SIE. However, it is inevitably oxidation of active O (Han and Oda, 2007). The Cl radical and O3 could
expensive to achieve complete mineralization of VOCs. react to form ClO radical, which could further react with trichloroeth­
The mass distribution of the chlorine element is shown in Fig. 5d. The ylene and its products (Vandenbroucke et al., 2014; Shang et al., 2021).
mass of the chlorine absorbed by the water sample increased from 43% C2 HCl3 + ⋅ OH → products
to 82% with increasing SIE. The majority of the chlorine element was
converted into Cl− in the water sample, while the remaining part of the k1 = 2.2 × 10− 12
cm3 s− 1
(10)
chlorine element was converted into organochlorine and Cl2. At the
same time, there were some yellow-green droplets attached to the inner C2 HCl3 + e → products
wall of the exhaust pipe. When these droplets were collected and placed
k2 = 2 × 10− 9 cm3 s− 1
(11)

C2 HCl3 + N ∗ → products

k3 = 2.2 × 10− 14
cm3 s− 1
(12)
( )
C2 HCl3 + O 3 P → products

k4 = 1.2 × 10− 13
cm3 s− 1
(13)

C2 HCl3 + O3 → products

k5 = 5 × 10− 20
cm3 s− 1
(14)

O ∗ + H2 O → ⋅ OH + ⋅OH

k6 = 2.2 × 10− 7 cm3 s− 1


(15)

⋅OH + ⋅ OH →H2 O2

k7 = 3 × 10− 11
cm3 s− 1
(16)

O3 + H2 O2 → ⋅ OH + H2 O ⋅ + O2

k8 = 5.5 × 106 cm3 s− 1


(17)
At the same time, nitrogen-containing polychlorinated organics were
also detected, which had not been reported in other articles. Previous
studies have confirmed the generation of active oxygen species that
Fig. 5. Proposed reaction pathways for the degradation of TCE. could oxidize TCE and intermediates (Song et al., 2019). However, Jiang

6
X.-R. Hu et al. Chemosphere 329 (2023) 138651

et al. (2022) used the optical emission spectra (OES) technique to reaction, and therefore the concentration of ozone slips (Tang et al.,
confirm that RNS was generated during the NTP processing of TCE. 2014). This also demonstrates why there is only NO2 and almost no
According to the byproducts detected (Fig. S5) and previous studies, a detected in the by-products.
proposed reaction scheme for the abatement of TCE in the atmosphere is
OH− + O3 → ⋅ O2 − + ⋅HO2
presented in Fig. 5. The removal of TCE depends on three mechanisms,
including: 1) Direct removal caused by the collision of high-energy
k9 = 3.1 × 106 cm3 s− 1
(18)
electrons with TCE; High-energy electrons would react with TCE to
produce Cl radicals and some chlorine-containing organic by-products; H + O3 = OH + O2
2) TCE molecules would react with free radicals to form
chlorine-containing organic products, which would then continue to k10 = 1.12 × 10− 10
exp( − 480/T ) cm3 s− 1
(19)
react with free radicals; and 3) The free radicals would completely
degrade the by-products to inorganic small molecules. HO2 + O3 = OH + O2 + O2

3.7. Formation of O3 under different conditions k11 = 1.0 × 10− 14


exp( − 490/T ) cm3 s− 1
(20)

Ozone is generally considered an unavoidable by-product of the air ⋅OH + O3 → ⋅ HO2 + O2


plasma reaction process. The concentration of ozone produced by DDBD
after treating air under different conditions is shown in Fig. 6. When the k12 = 1.9 × 10− 12
exp( − 1000/T) cm3 s− 1
(21)
RH of the inlet gas was maintained at about 60%, the ozone concen­
tration decreased with increasing inlet gas flow rates. At a flow rate of e + O2 → O + O + e
0.6 L min− 1, the ozone concentration reaches more than 1.4 g m− 3,
k13 = 4.2 × 10− 9 exp( − 5.6/T) cm3 s− 1
(22)
while the maximum ozone concentration was only 0.8 g m− 3 at a flow
rate of 3.2 L min− 1. The amount of ozone produced per unit of time was
O (3P) + O2 + N2 → O3 + N2
gradually increasing as the inlet gas flow rate increased. Because when
the inlet air flow rate increases, the amount of oxygen input per unit of
k14 = 6.2 × 10− 34
cm3 s− 1
(23)
time increases. The collision of high-energy electrons causes the disso­
ciation of oxygen molecules to produce oxygen atoms, which further ⋅NO + O3 → NO2 + O2
combine with oxygen molecules to produce ozone in air plasma (Jiang
et al., 2013). k15 = 1.5 × 10− 12
exp( − 1300/T) cm3 s− 1
(24)
Fig. 6b shows the effect of inlet RH on ozone concentration when the
inlet flow rate was 1.0 L min− 1. Ozone concentration decreased from a
maximum of nearly 2 g m− 3 to 1.2 g m− 3 with increasing inlet gas RH. 3.8. Biotoxicity analysis of by-products
From Eqs. 18–21, it can be concluded that ozone would react with the
discharge products of water molecules, such as OH− , ⋅OH, ⋅HO2 and Fig. 7a shows the change of SIE and the influence of substances with
other active particles, resulting in the depletion of ozone and oxygen different properties on luminescence inhibition. The sample that had not
atoms in wet air (Wang and Chen, 2009). Chen and Wang (2005) have been treated by DDBD (named “control” in Fig. 7), had no acute bio­
performed numerical simulations of ozone production using a DC corona toxicity. This is explained by the low solubility of TCE in water (water
discharge. The results showed that the ozone generation rate in dry air solubility at 25 ◦ C is 1280.94 mg L− 1). The acute biological toxicity of
was one order of magnitude higher than that in air with a relative hu­ the products and by-products gradually increased with the increase in
midity of about 50%, while the ozone generation rate only decreased SIE. The total acute toxicity to products increased from 0.75 mg L− 1
slightly in the relative humidity range of 50%–100%. Zn2+ to 1.16 mg L− 1 Zn2+ when the SIE increased from 64 J L− 1 to 202 J
It can be seen from Fig. 6 that the ozone concentration first increased L− 1. The sodium thiosulfate solution and the sodium hydroxide solution
and then slipped with the increase in the SIE. Because when the input were used as scavengers for ozone and acidic products, respectively,
power is changed to a lower range, Eq. (23) is dominant. As the SIE during the experiments. The results showed that the toxicity of the gases
increases, the temperature of the discharge space and the ozone con­ after DDBD treatment was mainly due to the production of toxic organic
centration increase to a certain value, and Eq. (22) and Eq. (24) begin to products. Moreover, the values of the produced acute biotoxicity and
dominate. The higher the SIE, the higher the temperature of the TOC trended close to each other. The acute biotoxicity of the organic
discharge space, the greater the reaction rate of the decomposition products first increased and then leveled off with increasing SIE, which

Fig. 6. Variation of ozone concentration with SIE under different conditions. (a. Ozone concentration at different inlet gas flow rates; b. Ozone concentration at
different RH.)

7
X.-R. Hu et al. Chemosphere 329 (2023) 138651

Fig. 7. a. Inhibition results of DDBD reactor exhaust gases (Absorption time t = 5 min); b. Acute toxicity to the oral rat of the TCE after treatment; c. Development
toxicity of the TCE after treatment.

was similar to the trend of the product TOC. At the same time, DDBD 4. Conclusions
generated high concentrations of ozone, which had a strong oxidizing
effect on luminescent bacteria, making the gas products more acutely In this study, the degradation of the TCE by non-thermal plasma
biotoxic. generated in a DDBD reactor was investigated. The experimental results
According to the above analysis, some organic by-products were of different parameters for TCE degradation indicated that the removal
generated during the degradation of TCE by DDBD. The Toxicity Esti­ efficiency was significantly enhanced with rising residence time or
mation Software Tool (T.E.S.T.) was applied to predict the quantitative reducing inlet gas concentrations, but higher residence time or lower
structure-activity relationship (QSAR) of these intermediate by-products inlet gas concentrations result in lower energy efficiency. Mass spec­
in this study (Wu et al., 2019). Fig. 7b–c shows the changes in the trometry analysis showed that most products of TCE were soluble in
toxicity of these by-products. As shown in Fig. 7b, the oral rat LD50 of DDBD, including trichloroacetic acid, trichloroacetaldehyde, dichlor­
TCE and dichloroacetyl chloride were 4915.06 mg kg− 1 and 2457.19 oacetyl chloride, dichloromethane, and other polychlorinated organic
mg kg− 1, respectively, which were deemed to be “toxicity”. The oral rat compounds. Due to dechlorination, the pH of the water absorption so­
LD50 of trichloromethane (2), dichloronitromethane (3), tri­ lution is significantly reduced. The concentration of O3 is reduced with
chloronitromethane (4), pentachloroethane (6), and tri­ the increase in inlet gas flow or relative humidity of the plasma device.
chloroacetaldehyde were 694.86 mg kg− 1, 256.19 mg kg− 1, 249.93 mg The O3 concentration decreased due to the increase in reactor temper­
kg− 1, 920.35 mg kg− 1, 464.03 mg kg− 1, respectively, which were sup­ ature under high SIE conditions. After DDBD treatment, the toxicity of
posed to be “very toxic”. Compared with TCE, the values of the oral rat the exhaust gas was significantly higher, which was attributed to the
LD50 of these byproducts all decreased. These results indicated that the conversion of TCE into more toxic products and the production of ozone
acute toxicity of intermediate by-products had increased. As shown in during the DDBD process. Overall, NTP technology should be adopted
Fig. 7c, the developmental toxicity of TCE was 0.37. The developmental carefully in isolation because of the potential generation of more toxic
toxicity of trichloromethane (2), dichloronitromethane (3), tri­ compounds.
chloronitromethane (4), pentachloroethane (6), trichloroacetaldehyde,
and dichloroacetyl chloride were 0.92, 0.76, 0.64, 0.81, 0.29, 0.92, and Author contribution
0.59, respectively. The result showed that the developmental toxicity of
intermediate by-products was also increased. In addition, Fig. 7d shows Xu-Rui Hu: Conceptualization, Methodology, Software, Writing –
that the mutagenicities of the intermediate byproducts are higher than original draft preparation. Yong-Chao Wang: Visualization and Editing.
those of TCE. Although phosgene (1), trichloromethane (2), and pen­ Zhen Tong: Experimental operation. Can Wang: Writing- Reviewing and
tachloroethane (6) have negative mutagenicities, the other byproducts Editing. Er-Hong Duan: Investigation. Meng-Fei Han: Software, Valida­
have positive mutagenicities. The comprehensive results show that the tion. Hsing-Cheng Hsi: Writing- Reviewing and Editing. Ji-Guang Deng:
residual toxicity of gases after the DDBD plasma treatment was signifi­ Supervision.
cantly increased, indicating that the TCE exhaust gas after the DDBD
treatment still needs to be treated. Declaration of competing interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence

8
X.-R. Hu et al. Chemosphere 329 (2023) 138651

the work reported in this paper. Li, S.J., Dang, X.Q., Yu, X., Yu, R., Abbasd, G., Zhang, Q., 2020. High energy efficient
degradation of toluene using a novel double dielectric barrier discharge reactor.
J. Hazard Mater. 400.
Data availability Li, Y.H., Cheng, S.W., Yuan, C.S., Lai, T.F., Hung, C.H., 2018. Removing volatile organic
compounds in cooking fume by nano-sized TiO2 photocatalytic reaction combined
Data will be made available on request. with ozone oxidation technique. Chemosphere 208, 808–817.
Li, Z., Luo, S.Q., Yang, Y., Chen, J.W., 2019. Highly efficient degradation of
trichloroethylene in groundwater based on peroxymonosulfate activation by
Acknowledgment bentonite supported Fe/Ni bimetallic nanoparticle. Chemosphere 216, 499–506.
Nguyen, H.P., Santos, C.A., Lee, T.J., Park, Y.K., Jo, Y.M., 2020. Decomposition of VOCs
using serial surface dielectric barrier discharge reactors. Environ. Technol. 43 (14),
This study was supported by the National Natural Science Founda­ 2145–2154.
tion of China (Grant No. 21961160743), Science and Technology Pro­ NIST, 2011. Chemical Kinetics Database. http://www.kinetics.nist.gov.
jects of Tianjin (21JCJQJC00080), Natural Science Foundation of Hebei Paz, Y., 2010. Application of TiO2 photocatalysis for air treatment: patents’ overview.
Appl. Catal., B 99, 448–460.
Province-Key Project (B2021208033). Santos, C.A., Phuong, N.H., Park, M.J., Kim, S.B., Jo, Y.M., 2020. Decomposition of
indoor VOC pollutants using non-thermal plasma with gas recycling. Kor. J. Chem.
Appendix A. Supplementary data Eng. 37 (1), 120–129.
Schiavon, M., Torretta, V., Casazza, A., Ragazzi, M., 2017. Non-thermal plasma as an
innovative option for the abatement of volatile organic compounds: a review. Water
Supplementary data to this article can be found online at https://doi. Air Soil Pollut. 228 (10), 388.
org/10.1016/j.chemosphere.2023.138651. Shahna, F.G., Bahrami, A., Alimohammadi, I., Yarahmadi, R., Jaleh, B., Gandomi, M.,
Ebrahimi, H., Abedi, K.A.D., 2017. Chlorobenzene degeradation by non-thermal
plasma combined with EG-TiO2/ZnO as a photocatalyst: effect of photocatalyst on
References CO2 selectivity and byproducts reduction. J. Hazard Mater. 324, 544–553.
Shang, K.F., Ren, J.Y., Zhang, Q., Lu, N., Jiang, N., Li, J., 2021. Successive treatment of
Alcantara-Garduno, M.E., Okuda, T., Nishijima, W., Okada, M., 2005. TCE removal from benzene and derived byproducts by a novel plasma catalysis-adsorption process.
porous media using an ozone-saturated solvent. Chem. Lett. 34 (11), 1504–1505. J. Environ. Chem. Eng. 9, 105767.
Chavadej, S., Kiatubolpaiboon, W., Rangsunvigit, P., Sreethawong, T., 2007. A combined Shen, Y.S., Ku, Y., 2002. Decomposition of gas-phase trichloroethene by the UV/TiO2
multistage corona discharge and catalytic system for gaseous benzene removal. process in the presence of ozone. Chemosphere 46 (1), 101–107.
J. Mol. Catal. Chem. 263 (1–2), 128–136. Song, H., Peng, Y.E., Liu, S., Bai, S.P., Hong, X.W., Li, J.H., 2019. The roles of various
Chen, J.H., Wang, P.X., 2005. Effect of relative humidity on electron distribution and plasma active species in toluene degradation by non-thermal plasma and plasma
ozone production by DC coronas in air. IEEE Trans. Plasma Sci. 33 (2), 808–812. catalysis. Plasma Chem. Plasma Process. 39 (6), 1469–1482.
Deng, J., Wu, F., Gao, S.X., Dionysiou, D.D., Huang, L.Z., 2022. Self-activated Ni(OH)2 Tang, X.J., Feng, F.D., Ye, L.L., Zhang, X.M., Huang, Y.F., Liu, Z., Yan, K.P., 2013.
cathode for complete electrochemical reduction of trichloroethylene to ethane in Removal of dilute VOCs in air by post-plasma catalysis over Ag-based composite
low-conductivity groundwater. Appl. Catal., B 309, 121258. oxide catalysts. Catal. Today 211, 39–43.
Dou, B.J., Bin, F., Wang, C., Jia, Q.Z., Li, J., 2013. Discharge characteristics and Tang, X.L., Gao, F.Y., Wang, J.G., Yi, H.H., Zhao, S.Z., Zhang, B.W., Zuo, Y.R., Wang, Z.
abatement of volatile organic compounds using plasma reactor packed with ceramic X., 2014. Comparative study between single- and double-dielectric barrier discharge
Raschig rings. J. Electrost. 71 (5), 939–944. reactor for nitric oxide removal. Ind. Eng. Chem. Res. 53 (14), 6197–6203.
Guo, T., Cheng, G.X., Tan, G.B., Xu, L., Huang, Z.X., Cheng, P., Zhou, Z., 2021. Real-time USEPA (United States Environmental Protection Agency), 2007. Toxicity and Exposure
analysis of intermediate products from non-thermal plasma degradation of ethyl Assessment for Children’s Health.
acetate in air using PTR-MS: performance evaluation and mechanism study. Vandenbroucke, A.M., Morent, R., De Geyter, N., Leys, C., 2011. Non-thermal plasmas
Chemosphere 264, 128430. for non-catalytic and catalytic VOC abatement. J. Hazard Mater. 195, 30–54.
Han, F.L., Li, M.Y., Zhong, H.R., Ting, L., Li, D.D., Shuo, Z., Guo, W.W., Fang, L., 2020. Vandenbroucke, A.M., Aerts, R., Van Gaens, W., De Geyter, N., Leys, C., Morent, R.,
Product analysis and mechanism of toluene degradation by low temperature plasma Bogaerts, A., 2014. Modeling and experimental study of trichloroethylene abatement
with single dielectric barrier discharge. J. Saudi Chem. Soc. 24 (9), 673–682. with a negative direct current corona discharge. Plasma Chem. Plasma Process. 35
Han, S.B., Oda, T., 2007. Decomposition mechanism of trichloroethylene based on by- (1), 217–230.
product distribution in the hybrid barrier discharge plasma process. Plasma Sources Wang, C., Xi, J.Y., Hu, H.Y., 2008. Chemical identification and acute biotoxicity
Sci. Technol. 16 (2), 413–421. assessment of gaseous chlorobenzene photodegradation products. Chemosphere 73
Jiang, L.Y., Zhu, R.Y., Mao, Y.B., Chen, J.M., Zhang, L., 2015. Conversion characteristics (8), 1167–1171.
and production evaluation of styrene/o-xylene mixtures removed by dbd Wang, C., Xi, J.Y., Hu, H.Y., Yao, Y., 2009. Advantages of combined UV
pretreatment. Int. J. Environ. Res. Publ. Health 12 (2), 1334–1350. photodegradation and biofiltration processes to treat gaseous chlorobenzene.
Jiang, N., Lu, N., Shang, K.F., Li, J., Wu, Y., 2013. Innovative approach for benzene J. Hazard Mater. 171 (1–3), 1120–1125.
degradation using hybrid surface/packed-bed discharge plasmas. Environ. Sci. Wang, P.X., Chen, J.H., 2009. Numerical modelling of ozone production in a wire-
Technol. 47 (17), 9898–9903. cylinder corona discharge and comparison with a wire-plate corona discharge.
Jiang, N., Kong, X.Q., Lu, X.L., Peng, B.F., Liu, Z.Y., Li, J., Shang, K.F., Lu, N., Wu, Y., J. Phys. D. 42 (3), 035202.
2022. Promoting streamer propagation, active species generation and Wang, Y.C., Lv, Y.H., Wang, C., Jiang, G.Y., Han, M.F., Deng, J.G., His, H.C., 2023.
trichloroethylene degradation using a three-electrode nanosecond pulsed sliding Microbial community evolution and functional trade-offs of biofilm in odor
DBD nanosecond plasma. J. Clean. Prod. 332, 129998. treatment biofilters. Water Res. 235, 119917.
Karatum, O., Deshusses, M.A., 2016. A comparative study of dilute VOCs treatment in a Wu, J.L., Xiong, Q., Liang, J.L., He, Q., Yang, D.X., Deng, R.Y., Chen, Y., 2019.
non-thermal plasma reactor. Chem. Eng. J. 294, 308–315. Degradation of benzotriazole by DBD plasma and peroxymonosulfate: mechanism,
Lei, C., Liang, F.Y., Li, J., Chen, W.Q., Huang, B.B., 2018. Electrochemical reductive degradation pathway and potential toxicity. Chem. Eng. J. 384.
dechlorination of chlorinated volatile organic compounds (Cl-VOCs): effects of Zhang, Z.X., Jiang, Z., Shangguan, W.F., 2016. Low-temperature catalysis for VOCs
molecular structure on the dehalogenation reactivity and mechanisms. Chem. Eng. J. removal in technology and application: a state-of-the-art review. Catal. Today 264,
358, 1054–1064. 270–278.

You might also like