Open access peer-reviewed chapter

A Novel Wastewater Treatment Method Using Electrical Pulsed Discharge Plasma over a Water Surface

Written By

Katsuyuki Takahashi, Koichi Takaki and Naoya Satta

Submitted: 31 July 2021 Reviewed: 04 November 2021 Published: 10 December 2021

DOI: 10.5772/intechopen.101494

From the Edited Volume

Sewage - Recent Advances, New Perspectives and Applications

Edited by Tao Zhang

Chapter metrics overview

437 Chapter Downloads

View Full Metrics

Abstract

Electrical pulsed discharge plasma produces various powerful oxidizing agents, such as hydroxyl radicals and ozone, which have high oxidation potential. These species play an important role in the decomposition of persistent organic compounds in wastewater. Because highly concentrated oxidants are directly produced inside the plasma, plasma realizes high-speed wastewater treatment without pretreatment of samples, such as pH adjustment. The pulsed discharge plasma generated over the water surface and inside bubbles is highlighted as a highly efficient method for plasma generation and radical supply into wastewater. In this paper, the physical and chemical properties of the discharge plasma generated over a water surface are described. The decomposition of persistent organic compounds dissolved in wastewater, such as 1,4-dioxane, formic acid, and dichloromethane, by plasma discharge is demonstrated, and their mechanisms are discussed. These persistent compounds, which have strong toxicity and stability, can be efficiently decomposed and removed quickly from solutions by plasma treatment. Furthermore, the treatment of nutrient solutions used in hydroponic systems for plant cultivation is also introduced as a novel application of plasma, and the effects of bacterial inactivation, decomposition of allelochemicals, and improvement in plant growth by plasma are demonstrated.

Keywords

  • plasma
  • wastewater
  • persistent compounds
  • advanced oxidation process
  • hydroxyl radical
  • oxidation
  • ozone

1. Introduction

Electrical pulsed discharge under water, an advanced oxidation process (AOP), can instantaneously produce chemically active species such as hydroxyl radicals and ozone in a nonthermal plasma [1, 2, 3]. It is well known that these species are powerful oxidizing agents that play an important role in degrading persistent organic chemical compounds and sterilizing bacteria in wastewater [4, 5]. Hydroxyl radicals can contribute to the decomposition of persistent organic pollutants, which cannot be decomposed by conventional methods because of their very high oxidation potential. Wastewater treatment using discharge plasma has several advantages over other AOPs. In the plasma system, active species are directly produced in plasma through various reactions initiated by high-energy electron impact. These active species are directly dissolved into water, and then oxidation reactions are induced. Because only electricity-supplied metal electrodes and a high voltage power supply are needed to generate plasma, pretreatment of water, such as pH adjustment is not required for active species production, unlike other AOPs such as Fenton and ozone/hydrogen peroxide methods [6]. Plasma can also be adapted for the treatment of water, which has ultraviolet (UV) light non-transmitting properties, which is advantageous over AOPs utilizing UV light, such as photocatalytic processes and UV-hydrogen peroxide methods. In the case of plasma discharge in contact with water, the types of active species produced in plasma can vary by atmospheric gas species. The production of active species can be localized near the plasma channel. Thus, water treatment using plasma discharge can be useful in various uses. Furthermore, because the density of active species in plasma is very high, it can also be used for the treatment of highly concentrated wastewater. These characteristics of discharge plasma can make the system simple, easy to use and easy to install anywhere. Owing to these advantages, the plasma technique has attracted attention as a promising method in various fields, such as industrial and environmental wastewater treatment [4, 6, 7] and material [8], agricultural [9, 10], and medical applications [11].

In this article, characteristics of discharge plasma generated by pulsed high voltages, chemical reactions induced by plasma, and treatment of persistent organic pollutants such as 1,4-dioxane and dichloromethane dissolved in water by plasma discharges are introduced. Agricultural applications of discharge plasma in contact with water surfaces are also mentioned as a novel application.

Advertisement

2. Generation of pulsed discharge plasma

Plasma is one of the states of matter and consists of high-density electrons and ions produced through ionization reactions of neutral gas atoms and molecules. Ionization is initiated by an impact of high-energy electrons accelerated by a high electric field; then, the electron and ion densities rapidly increase due to an electron avalanche phenomenon, and plasma forms [12]. To generate a high electric field, a sharpened metal electrode to which a high voltage is applied is generally used. Although various kinds of voltage waveforms, such as DC and AC, are employed, a pulsed high voltage generated by a pulsed power generator is used for highly efficient plasma generation, [13, 14] especially under water. Pulsed power is a special technology used to generate a pulse voltage that has a high voltage amplitude over several tens of kV, high power over several tens of kW, short pulses on the order of nanoseconds, and high pulse frequencies over several kHz, achieved with a compact system [15].

When a pulsed high voltage is applied to pointed electrodes immersed in water, discharge plasma is generated at the tip of the electrode. One of the formation mechanisms of the discharge plasma in water using a long pulse voltage on the order of several μs is considered as follows: In the first step, the discharge is initiated in a small bubble on the electrode surface, and then it propagates into the water, as shown in Figure 1. The current in the high electric field region causes Joule heating and vaporization of the liquid, forming bubbles [16, 17, 18]. The electron density in the plasma directly generated under water is very high, on the order of 1025 m−3 [19], which can produce very high density chemically active species such as hydroxyl radicals and contribute to high-speed wastewater treatment. The plasma discharge under water can also produce strong shock waves with pressure on the order of several GPa. The problem with this method is that wastewaters exhibit high conductivity due to the high concentration of various substances. When a voltage is applied to the electrodes immersed in water, which has high conductivity, a large current flow, and large energy are dissipated during Joule heating, which does not contribute to chemically active species production. The discharge volume decreases with increasing conductivity of the water at the same input energy. As a result, the energy efficiency for water treatment by discharge decreases to a low value under high-conductivity water conditions. Furthermore, because the breakdown strength of water is 300 MV/m and is 10 times higher than that of air (3.2 MV/m), high voltage is required to generate plasma discharges under water and to increase plasma volume. To solve this problem, methods to generate a pulsed discharge in contact with water using an electrode in the gas phase have been investigated [20, 21, 22]. Because the electrode is insulated by gas, the energy loss caused by the conductive current is significantly reduced. Plasma generation in the gas phase is much easier than that in the liquid phase, which contributes to a decrease in operating voltage and makes the system compact and light.

Figure 1.

Schematic illustration of discharge plasma using electrodes immersed in water.

Various methods to generate a pulsed discharge in contact with water have been proposed [23, 24]. Figure 2 shows schematic illustrations and photographs of examples. The simplest method is to generate discharge that propagates and spreads over the water surface by using a high-voltage electrode placed above stationary water, [25, 26] as shown in Figure 2(a). The chemical species generated in the plasma discharge are dissolved into the water and react with organic compounds in the water as described later. The discharges generated inside bubbles as shown in Figure 2(b), which are produced by injecting gas into the water using a gas feed tube, have been widely investigated [23, 26, 27, 28]. This method has advantages such as a high ratio of water surface to gas volume, easy control of the gas purity and components, easy use in various fields, and automatic water circulation. The reactor can be simply constructed as shown in Figure 3(a) [29]. This reactor consists of a sealed glass vial for gas chromatography and glass tubes in which wire electrodes are inserted. Because the glass vial is sealed, the purity of the gas component in the reaction area in the vial can be improved. The treatment speed and volume can be increased easily by increasing the number of gas feed tubes and electrodes, as shown in Figure 3(b). Figure 3(c) shows a reactor consisting of a separator film that has tiny pores and a high voltage wire electrode in the gas phase [20]. Multiple discharge plasmas occur simultaneously on the wire electrode surface, which propagates into the bubbles. The separator film is used to separate the liquid phase and gas phase for insulation of the high voltage electrode. Instead of the separator and the wire electrode, ceramics with millipore needle electrodes have been used in some studies [7, 30].

Figure 2.

Schematic illustrations and photographs of pulsed discharge in contact with water generated (a) above the stationary water surface, (b) inside the bubble, and (c) area where droplets are sprayed.

Figure 3.

A reactor used for pulsed discharge generation inside bubbles using (a) a single electrode inserted in a sealed glass vial, (b) parallel electrodes inserted in a vessel, and (c) a separator with multiple pores and a high voltage wire electrode placed in the gas phase.

When a pulsed high voltage is applied to the high-voltage electrode in the gas phase, the discharge propagates into the bubble surface from the high-voltage electrode [31, 32]. Figure 4 shows typical voltage–current waveforms and flaming photographs of the pulsed discharge generated in an argon bubble, as shown in Figure 2(b) taken by using an intensified CCD camera with an exposure time of 5 ns. At time t0, a corona discharge occurs at the tip of the electrode, and then, the discharge propagates to the tip of the glass tube along the surface of the glass at t1 and t2. Inside the glass tube, discharge occurs at not only the tip of the electrode but also the side of the electrode, as observed with dielectric discharges. At t3, the discharge propagates along the bubble surface and does not travel to the bubble at t3 to t5. The results obtained by numerical simulation also show that a large part of the discharge inside the bubble propagates along with the bubble with similar size of bubbles [33, 34]. The propagation velocity of the discharge is 2.7–3.6 × 105 m/s in the Ar gas bubble [32] and 5 × 105 m/s on the surface of a glass tube filled with air, [31] which correspond to the surface discharge over a dielectric material under atmospheric pressure [35, 36]. When the voltage drops, the discharge disappears at t6, but back discharge occurs again at t7 due to the electric field formed by charges accumulated over the bubble surface.

Figure 4.

(a) Typical voltage–current waveforms and (b) flaming photograph of discharge generated inside argon bubble.

Conductivity is the most important parameter for discharge propagation. Generally, the energy efficiency for discharge generation and propagation over a water surface is strongly affected by the conductivity of the water [32, 37]. Figure 5 shows the equivalent circuit model [38] of (a) discharge in contact with water using the electrode system shown in Figure 2(b) discharge generated under water using the electrode system shown in Figure 1. When a breakdown occurs, switch SW1 is turned on. The plasma impedance ZP(t) depends on the electron density of the plasma and changes with time during voltage application. The equivalent circuit of the water surface is expressed as the resistor RL and capacitor CL connected in parallel. The resistance of the resistor decreases with increasing conductivity. The discharge in contact with water propagates with the charging water surface. The decrease in resistance enhances the leakage of the accumulated charge on the surface of the bubble with decreasing conduction relaxation time, which is expressed by dividing the permittivity by the conductivity of the solution, which concentrates the discharge current and decreases the electric field. Therefore, the discharge length decreases with increasing conductivity. For example, in the case of pulsed discharge generated inside a bubble as shown in Figure 4, the maximum discharge length decreases by 50% by increasing conductivity from 7 μS/cm to 1000 μS/cm [32].

Figure 5.

Equivalent circuit model of (a) plasma contacted in water and (b) plasma generated under water.

When the electrode is placed in the gas phase and is not in contact with water, conductive current does not flow into the water without plasma generation, i.e., without SW1 turning on. On the other hand, when the electrode is placed in water, conductive current always flows into the water, which causes energy loss, and the energy efficiency for active species production decreases. This is one of the reasons that the energy efficiency of a wastewater treatment method using plasma contacted water is much higher than that using plasma directly generated under water [39]. Using this method, the energy efficiency can be improved over tens of times, especially for highly conductive wastewater treatment [39].

To reduce the influence of the conductivity, a method to generate streamer discharges into the area where droplets are sprayed using a pulsed high voltage is proposed as shown in Figure 2(b) [40]. This is one of the most efficient methods to supply chemical species into the water because streamer generation in the gas phase is not affected by water droplets, and the contact area between the discharge and water is large.

Advertisement

3. Production of chemical species by discharge plasma

In the plasma generated in the gas phase, various chemically active species such as ozone (O3), oxygen radicals (O), and hydroxyl radicals (OH) are produced and dissolved at the water surface. Since these species have a high oxidation potential, they can contribute to wastewater treatment, i.e., the decomposition of organic compounds and the inactivation of bacteria. The production reactions are initiated by the impacts of high-energy electrons on neutral molecules. Representative reactions in a gas that contains O2, N2, and Ar are as follows: [41, 42, 43].

O2+e>8.4eVO1D+O3P+eE1
O2+e>5.58eVO3P+O3P+eE2
O3P+O2+MO3+ME3
H2O+O1D2OHE4
N2+e>9.0eVN4S+N2D+eE5
N2D+H2OOH+NHE6
N2D+O2NO+O1DE7
N2D+O2NO+O3PE8
Ar+e>11.55eVAr3P+eE9
Ar3P+H2OAr+OH+HE10
H2O+e>7.1eVOH+H+eE11
H2O+e>12.62eVH2O++2eE12
H2O++H2OH3O++OHE13

where (3) is a three-body recombination reaction and a neutral molecule such as O2, N2 or Ar is the third collision partner (M), which takes part in energy absorption, but does not react chemically. Ar, a noble gas, is generally used for highly efficient plasma generation at low cost. Because the reaction rates of (11) and (12) are not high in atmospheric nonthermal plasma, OH is mainly produced by indirect reactions such as reactions (4), (6) and (10) [44, 45, 46]. When a water surface is used as the cathode, OH is also produced under water by the bombardment of high-energy positive ions produced inside plasma, such as H2O+ and Ar+, to the water surface [47, 48].

Ar+e>15.8eVAr++2eE14
Ar++H2OH2O++ArE15
H2O+OH+H+E16

Figure 6 shows typical emission spectra of the discharge inside the Ar or O2 bubble. The peaks at wavelengths of 282 and 308 nm are the A2Σ+ (v = 1) → X2Π (v = 0), and A2Σ+ (v = 0) → X2Π (v = 0) emission lines of OH, respectively. The peak at 656 nm is attributed to a Balmer H-alpha emission [32]. These peaks are strongly observed in argon gas because of its high electron density. In O2 gas, the peak at a wavelength of 774 nm corresponds to the atomic oxygen (O) transition of 3p5P-3s5S0. Figure 7 shows the photograph of discharge inside an Ar bubble taken by a high-speed intensified CCD camera with and without a bandpass filter with a bandwidth of 310 ± 10 nm, showing light emission with a wavelength corresponding to an excitation line of OH (308 nm). The photograph shows that the excitation-emission of OH with a wavelength of 308 nm is observed at the same position of the plasma channel generated along the bubble surface observed without the bandpass filter. This result shows that OH is produced inside the plasma channel.

Figure 6.

Typical emission spectra of discharge inside argon or oxygen bubble.

Figure 7.

Photograph of discharge inside the bubble (a) without and (b) with a bandpass filter (310 ± 10 nm).

Because these oxidants have high reactivities, they are consumed in loss reactions in both the gas and aqueous phases, which decreases the efficiency of wastewater treatment as shown in Figure 8. OH is well known as the most powerful oxidant and has a very short lifetime. OH is mainly lost by three-body recombination reactions, as shown in reaction (17) in the gas phase and reactions (18) and (19) in the aqueous phase [49, 50].

Figure 8.

Schematic illustrations of reactions of OH produced by plasma at the vicinity of water surface.

OH+OH+MH2O2E17
OH+OHH2O2E18
OH)+HH2OE19

The lifetime and diffusion constant of OH in the gas phase are on the order of 10−5 s and 10−10 m2/s, respectively, and the diffusion length is several tens of μm [51]. The lifetime of dissolved OH is on the order of 10−6 ∼ 10−7 s with penetration lengths on the order of 10−5 ∼ 10−6 m [52, 53, 54]. Figure 9 shows the hydrogen peroxide (H2O2) concentration produced by the OH recombination reaction, as described later, as a function of treatment time by plasma generated inside Ar or O2 bubbles. In this case, 100 mL of dye solution was treated by plasma discharges generated inside bubbles produced by multiple pores, as shown in Figure 3(c). The gas flow rate is 4.5 L/min. Pulse voltages with an amplitude of 20 kV, a pulse width of 160 ns, and a pulse repetition rate of 250 Hz are applied to the wire electrode placed in the gas phase to generate plasma [73]. The electric power into the plasma is the same in both gas species at 3.4 W. H2O2 concentration, i.e., OH production, with Ar gas is higher than that with O2 gas, which is in good agreement with the emission spectra, as shown in Figure 6. Thus, Ar injection is preferred for highly efficient OH production.

Figure 9.

Time change of the H2O2 concentration produced by plasma treatment.

Many researchers have analyzed the amount of dissolved OH in water using chemical probes such as terephthalic acid [55, 56]. The energy yield for the surface pulsed streamer discharge on the liquid with argon gas is estimated to be 13 ∼ 27 × 10−9 mol/J with disodium terephthalate solution [57]. The typical energy yield of dissolved OH in water supplied by plasma, which can react with solutes in water, is estimated to be 4 ∼ 5 × 10−9 mol/J using the indigo carmine solution as a chemical probe [29]. Because plasma directly generated under water has a high OH density, the reaction rate of OH loss reactions is very high, which decreases the energy efficiency for OH supply.

When the feeding gas contains O2, highly concentrated O3 is generated in the plasma through reactions (1) ∼ (3). Henry’s law constants for O3 and OH are 0.011 and 30 M/atm at 298 K, [58] respectively, which shows that the solubility of O3 is lower than that of OH. Additionally, the oxidation potential of O3 is lower than that of OH. However, the lifetime of O3 in water is on the order of several min [59, 60, 61] and much longer than that of OH. Thus, when wastewater contains only organic compounds that can be easily oxidized by O3, the treatment efficiency utilizing O3 is higher than that utilizing OH. In this case, O2 is preferred for use as a feeding gas for high-efficient O3 production. O3 dissolved in water can contribute to OH production via the following reaction with hydroperoxide anion (HO2) species, the conjugate base of H2O2, under alkaline conditions with a high reaction rate [62].

O3+H2O2O2+OH+HO2E20
H2O2HO2+H+pKa=11.6E21
O3+HO2O2+OH+O2E22

The pKa value of the equilibrium reaction (R14) is 11.6. Because the reaction rate of O3 and H2O2 is very low (10−2 M−1 s−1), O3 mainly reacts with HO2 with a reaction rate constant of 5.5 × 106 M−1 s−1 and hardly reacts with H2O2 directly. Thus, pH adjustment is required to induce this OH production reaction.

When the feeding gas contains nitrogen, nitrous (HNO2) and nitrite (HNO3) acids, are generated by plasma as byproducts through the following reactions; [46, 63, 64, 65].

O3P+NO+N2NO2+N2E23
O3+NONO2+O2E24
NH+OHNO+H2E25
OH+NO+MHNO2E26
OH+NO2HNO3E27
3HNO2HNO3+2NO+H2OE28
HO2+NO2O2NOOHE29
H2O2+HNO2HOONO+H2OE30
HOONO+H2O2O2NOOH+H2OE31

Peroxynitric acid (HOONO2) contributes to HOO production via equilibrium decomposition reaction, [66].

HOONO2HOO+NO2E32

When HNO2 and HNO3 are dissolved in water, nitric and nitrate ions are easily produced in water through acid dissociations of HNO2 and HNO3 because of their low pKa values, of 3.35 and − 1.8, respectively, and the pH of the solution significantly decreases [65, 67].

HNO2H++NO2E33
HNO3H++NO3E34

NO2 is oxidized to NO3 by O3 in solution [68]. When the pH is lower than 3.5, NO2 is oxidized by H+. These reactions are follows:

NO2+O3NO3+O2E35
NO2+3H+NO+NO3+H3OorH2O+H+E36

These nitrogen species can contribute to the inactivation of bacteria in wastewater as described later. In the case of plasma discharges inside multiple air bubbles using a reactor similar to that in Figure 3(b) with eight glass tubes in which total air gas with a flow rate of 5 L/min is injected, approximately 9 mM NO3 is generated in 20 L of solution with 24 hours of plasma treatment, [69] which is a sufficient concentration for using fertilizers in plant cultivation. This effect is described in Section 5.

When metals that are easily oxidized by electrolysis, such as iron and copper, are used as an electrode immersed in water, metal ions such as Fe2+, Fe3+, Cu+, and Cu2+ are generated by electrooxidation by a large discharge current, and the metal ion concentration increases. These metal ions contribute to the oxidation of organic compounds in wastewater not only directly but also through OH production by redox reactions well known as Fenton reactions with H2O2 produced by plasma as follows: [70, 71, 72, 73].

Fe2++H2O2OH+OH+Fe3+E37
Fe2++OHFe3++OHE38
Fe3++H2O2Fe2++HO2+H+E39
Fe3++HO2Fe2++O2+H+E40
Fe2++HO2Fe3++HO2E41
Fe3++eFe2+E42
H+Fe3+Fe2++H+E43
Cu++OHCu2++OHE44
O2+O2+2H+H2O2E45

Reactions related to iron ions are well known as Fenton reactions and occur under acidic conditions. Generally, electrodes immersed in water are used as grounded electrodes to keep the voltage potential of water zero for safety reasons. Therefore, when a positive (or negative) high voltage is applied to a high voltage electrode, iron (or copper) ions are dissolved from the grounded electrode immersed in water and used as the anode (or cathode). For example, in the case of 1,4-decomposition as described in Section 4, the decomposition efficiency can be improved by approximately three times, by using iron wire as an electrode immersed in water, compared with that achieved using stainless steel wire [73].

The chemical species dissolved in the water react with not only organic compounds but also inorganic compounds. Table 1 shows reaction the rate constants for O3 and OH reacting with inorganic ions [74, 75, 76, 77, 78]. Halogen ions such as chloride ions (Cl), bromide ions (Br) and iodide ions (I) have high reactivity toward O3 and OH and act as O3 and OH scavengers, which inhibits the decomposition reactions of organic compounds. Furthermore, toxic compounds such as hypochlorite ions (ClO) and bromate ions (BrO3) are produced via oxidation reactions, as shown in Figures 10 and 11. Carbon dioxide radicals (CO2), carbonate radicals (CO3), and sulfate radicals (SO4) are produced through reactions of OH with formate ions (HCO2), carbonate ions (CO32−), and hydrogen sulfate ions (HSO4) with hydroxyl radicals in liquid [79]. These radicals also contribute to the decomposition of organic compounds dissolved in wastewater as synergistic effects. Although hydroxide ions (OH) and per hydroxyl ions (HO2) act as scavengers of OH, OH can be produced via reaction with O3 under high-pH conditions. Thus, the influence of these ions on treatment efficiency and water quality after wastewater treatment using plasma should be considered.

SoluteReaction rate constant for O3 [M−1 s−1]Reaction rate constant for OH [M−1 s−1]
Cl0.0034.3 × 109
Br1601.1 × 109
I1.2 × 1051.1 × 1010
NO23.7 × 1051.0 × 1010
NO31.0 × 105
HSO33.7 × 1054.5 × 109
SO32−1.5 × 1095.1 × 109
HSO4<10−46.9 × 105
H2SO41.4 × 107
H2O20.012.7 × 107
HO25.5 × 1067.5 × 109
HCO30.018.5 × 106
HCO21005.1 × 109
CO32−0.013.9 × 108
CO21 × 106
OH2101.2 × 1010

Table 1.

Reaction rate constant for O3 and OH reacting with inorganic compounds.

Figure 10.

Schematic illustrations of reactions involving Cl.

Figure 11.

Schematic illustrations of reactions involving Br.

Advertisement

4. Decomposition of persistent organic pollutants

Organic pollutants entering the water environment cause industrial, agricultural, domestic, livestock, and medical wastewater drainage into and leakage from polluted soil. Because most organic pollutants have strong toxicity, carcinogenicity, and genotoxicity, even at low concentrations, it is urgent to develop effective technologies for the highly efficient removal of these contaminants from water. 1,4-Dioxane has been widely used as a solvent. Because it has high solubility and is hard to absorb to the bottom material, it can be widely spread in water environments [80, 81]. Dichloromethane (DCM) is a chlorinated volatile organic compound (Cl-VOC) and has high volatility. It is also used as a solvent, cleaner, degreasing agent, and liquid fuel and released into the environment through water discharge from chemical industries [82, 83]. These compounds are very difficult to decompose using conventional oxidation treatment methods such as chlorine, ozone oxidation, and bioremediation owing to their chemical and biological stability.

In wastewater treatment systems using plasma discharges, high-density powerful oxidants such as OH and O3 produced by plasma contribute to the decomposition of persistent organic pollutants and the inactivation of bacteria. This method provides high-speed treatment and decomposition of organic compounds that are difficult to remove with conventional methods and has attracted much attention as an advanced oxidation technology. Many researchers have proposed that plasma treatment is effective for the decomposition of various kinds of toxic aromatic hydrocarbons, such as chlorophenol, [84] naphthalene, toluene [85], and bisphenol A [86], and volatile organic compounds such as trichloroacetic acid [87]. Oxidants also contribute to the quick decolorization of wastewater containing dye solutions [88, 89]. In this section, the decomposition of various kinds of organic compounds in water is systematically described. The decomposition characteristics of indigo carmine, a common blue dye, 1,4-dioxane [73] which has a high solubility in water, and formic acid, [90] which is an intermediate product of various organic compounds, such as dichloromethane (DCM), [91] which is a volatile organic compound, are introduced as an example.

Chemically active species dissolve in water in the vicinity of the water surface and induce oxidation reactions such as hydrogen atom abstraction, electron transfer, electrophilic addition, and chain oxidation reactions triggered by these reactions. The following reactions are typical oxidation reactions of organic compounds induced by OH.

RH+OHR+H2OE46
R+OHR++OHE47
R+OHR+OHE48
R+OHROHE49
OH+R2C=CR2R2OHCCR2E50
R+O2RO2E51
R+RRRE52
R+RO2RO2RE53
RO2+RHROOH+RE54
RO+RHROH+RE55
RO2+HO2RO+O2+OHE56
RO2+RO2RO2RE57
E58

Indigo carmine and azo phloxine are common dyes and have blue and red colors, respectively. Figure 12 shows the time change of appearance and UV–vis spectra of two dye solutions, (a) indigo carmine and (b) azo phloxine, by treatment using plasma generated inside O2 gas bubbles. In this case, 100 mL of dye solution was treated by plasma discharges generated inside bubbles produced by multiple pores as shown in Figure 3(c). The gas flow rate is 4.5 L/min. Pulse voltages with an amplitude of 20 kV, a pulse width of 160 ns, and a pulse repetition rate of 250 Hz are applied to the wire electrode placed in the gas phase to generate plasma [73]. The energy input into a pulse is fixed at 13.5 mJ for each gas species by adjusting the voltage amplitude. From the input energy and pulse repetition rate, the input power can be estimated as 3.4 W. Indigo carmine and azo phloxine have absorbance peaks at wavelengths of 610 nm and 531 nm, which are responsible for the blue and red colors of their solutions, respectively. These peaks decrease, the colors disappear after 20 s and 180 s of plasma treatment, and then the solutions are fully decolorized after treatment. This result shows that plasma treatment is effective for the decolorization treatment of wastewater. Figure 13 shows the decolorization rate of these solutions using Ar and O2 as feeding gases to generate bubbles. Decolorization of both solutions in the case of O2 injection is much faster than that of Ar injection. The energy efficiency for 50% decolorization of indigo carmine is 1.3 × 103 mmol/kWh for O2 injection and 6.1 mmol/kWh for Ar injection, and those of azo phloxine are 77 mmol/kWh for O2 injection and 4.8 mmol/kWh for Ar injection. The absorbance peaks of these dyes are due to the carbon–carbon double bond in the H-chromophore of indigo carmine and the nitrogen-nitrogen double bond of the azo chromophore of azo phloxine as shown in Figure 14, which are responsible for the blue and red colors of the solution. These double bonds have high electron density and high reactivity to electrophilic OH. It is well known that O3 reacts with C-C and N-N double bonds rapidly; however, the reaction of OH with dyes is at least four orders of magnitude faster than that of O3. As already mentioned, the lifetime of OH in water is much shorter than that of O3 because of its high reactivity and high rate constant of recombination loss reactions, and only OH generated in the vicinity of the bubble surface can be dissolved in the solution. In this case, these two dyes can be easily oxidized by O3. Therefore, O2 injection to generate O3 at high efficiency is preferred for the decomposition of compounds that have high reactivity to O3. The reactivity of an azo group with ozone is very low compared to that of olefins, [92] which is one of the reasons that indigo carmine is more readily decolorized than azo phroxine.

Figure 12.

Time change of appearance and UV–vis spectra of two dye solutions, (a) indigo carmine and (b) azo phloxine, by treatment using plasma generated inside O2 gas bubble.

Figure 13.

Time change of decolorization rate of indigo carmine and azo phloxine solutions by plasma treatment using Ar and O2 gases.

Figure 14.

Chemical structure of (a) indigo carmine and (b) azo phloxine.

Since 1,4-dioxane has low reactivity to O3 due to its high chemical stability, plasma treatment is a promising method for 1,4-dioxane removal. Figure 15 shows one of the proposed decomposition processes of 1,4-dioxane by hydroxyl radicals [93, 94, 95]. The reaction is initiated by hydrogen atom abstraction, and various reactions, such as chain reactions, are involved. Carboxylic acids such as formic acids (HCOOH) and oxalic acid ((COOH)2) are produced as intermediate products. Figure 16 shows the time change of the TOC concentration of the 1,4-dioxane solution with an initial concentration of 0.23 mM with two different injected gases and with and without discharges. The experimental conditions were the same as those mentioned above. The electric power in the plasma is the same in both gas species. The TOC decreases without discharges because of the volatilization by bubbling gas, and the TOC concentration in the O2 injection is almost the same without discharges. The rate of TOC decrease rate in the case of Ar injection is much higher than that in the case of O2 injection. 1,4-dioxane hardly reacts with O3 because of its low reaction rate (0.32 M−1 s−1) [78]. The rate constant of reaction between OH and 1,4-dioxane is 2.4*109 M−1 s−1 [96]. The production rate of OH in the case of Ar injection is much higher than that in the case of O2 injection. Thus, Ar injection is preferred for the decomposition of organic compounds that have low reactivity to O3. The energy efficiency for 1,4-dioxane decomposition is 2 mmol/kWh.

Figure 15.

Oxidation process of 1,4-dioxane by OH.

Figure 16.

Time change of the TOC concentration of a 1,4-dioxane solution by plasma treatment.

Figure 17 shows the oxidation of acetic acid (CH3COOH) by OH radicals. CH3COOH has also low reactivity to O3. It can be decomposed into formic acid, oxalic acid, and finally, CO2 and H2O by oxidation reactions with OH. These acids have an acidity constant and are in equilibrium with their conjugated bases. For example, formic acid is in equilibrium with its conjugate base, formate, as follows:

Figure 17.

Oxidation process of CH3COOH by OH.

HCOOHHCOO+H+E59

The pKa value of this equilibrium reaction is 3.75. The reaction processes of formic acid and formate with OH and O3 are represented by the following: [76, 78].

HCOOH+O3CO2+OH+HO2E60
HCOOH+·OHH2O+COOHE61
HCOO+O3CO2+OH+O2E62
HCOO+OHH2O+COOE63

The rate constant of reaction (62) is 100 M−1 s−1 and is 20 times higher than that of reaction (60) (5 M−1 s−1). Similarly, the rate constant of reaction (63) is 3.2 × 109 M−1 s−1 and is 25 times higher than that of reaction (61) (1.3 × 108 M−1 s−1). Therefore, decomposition of these acids at higher pH is preferred. It has been reported that the decomposition rate of formic acid can be improved twice by increasing the pH from 3.5 to 4.2 with the addition of sodium carbonate, in the case of plasma generated inside oxygen bubbles using the reactor shown in Figure 3(a) [90]. Thus, pH adjustment is an important factor for wastewater treatment, while active species can be directly supplied from plasma.

The decomposition process of DCM is different from that of other soluble compounds because of its high volatility. The rate constant of DCM reacting with OH in a liquid is 2.2 × 107 M−1S−1 [97] and is much lower than that of reactions of other soluble organic compounds, such as 1,4-dioxane, phenol, and formic acids, with OH (on the order of 109 M−1 s−1). Figure 18 shows the TOC and the amount of decomposed DCM as a function of treatment time using plasma generated inside Ar bubbles using the reaction shown in Figure 3(a). The gas flow rate is 30 mL/min. Pulse voltages with an amplitude of 16 kV, a pulse width of 160 ns, and a pulse repetition rate of 250 Hz are applied to the wire electrode placed in the gas phase to generate plasma [91]. The amount of decomposed DCM is calculated from the Cl concentration detached from DCM during the decomposition process. The TOC decreases without discharges because of volatilization by bubbling gas. DCM can be successfully decomposed by discharges generated inside Ar bubbles [91]. The energy efficiency for DCM decomposition is approximately 60 mmol/kWh and is much higher than that for decolorization of dye solutions and 1,4-dioxane decomposition, as mentioned above, although the rate constant for DCM reacting with OH is low. This result suggests that almost all DCM is decomposed in the gas phase in the bubble, not in the liquid phase. DCM is quickly volatilized due to its high volatility and is transferred into the gas phase in the bubble when Ar gas is fed into the DCM solution, and DCM is exposed by plasma discharges generated inside the bubble. DCM can be decomposed by OH with the following reactions: [98].

Figure 18.

Time change of the TOC concentration of DCM solution and amount of decomposed DCM by plasma treatment.

CH2Cl2g+OHgCHCl2g+H2OgE64
CHCl2g+OHgCHOClg+HClgE65
CHOClg+OHgCOClg+H2OgE66
COClg+OHgCO2g+HClgE67

The Cl-H bonding energy of CH2Cl2 is 4.28 eV and is lower than the energy of metastable Ar (11.55 eV) produced by electron impact, as shown in reaction (9) [99]. Therefore, CH2Cl2 is also possibly decomposed by Penning ionization by the energy transfer collision of metastable Ar (Ar*) as shown in the following reactions: [99, 100].

CH2Cl2g+ArgCH2Cl2+g+ArgE68
CH2Cl2+gCH2Cl+g+ClgE69
CH2Cl+g+egCH2Clg+ClgE70

Because the concentration of DCM inside the gas bubble is on the order of several hundred ppm as mentioned previously, and is much lower than the Ar and H2O concentrations, the DCM dissociation by electron impact is negligible [98, 101].

Advertisement

5. Novel applications for agriculture

Agricultural applications of plasma have been widely investigated in recent years and have become one of the most attractive research topics in plasma science [102, 103, 104, 105]. As already mentioned in this article, plasma in contact with water can produce various types of chemically active species, which have various effects on the environment of plant growth.

As a production process of fruits and vegetables, hydroponics, which is the method of growing plants without soil using the nutrient solution, has been widely used. In hydroponics, the nutrient solution is recirculated in a closed system to reduce the cost and the environmental load. In the system, plant diseases caused by microbial contamination of artificial nutrient solution rapidly spread in the circulation system and cause serious damage to the entire plant. During the entire period of plant growth, contamination with pathogens can never be excluded since pathogens are introduced in the nutrient solution via the irrigation water supply. Therefore, the nutrient solution should be remedied by continuous water treatment system operation during the cultivation period. Active species such as OH and O3 produced by discharges can contribute to the inactivation of pathogenic bacteria-contaminated in liquid fertilizer. A plasma treatment system for the inactivation of Ralstonia solanacearum bacteria in liquid fertilizer for a tomato hydroponic culture system has been developed [106]. The number of colony-forming units (CFU) of R. solanacearum in the liquid fertilizer decreased from 107 to 102 CFU/mL when treated with discharge plasma. Tomato seedlings treated with discharge plasma were relatively healthy, while the infected positive controls all wilted and died, as shown in Figure 19. Plasma was also used for the removal of allelochemicals, organic compounds exuded from the root of plants, which have autotoxic effects on plant growth. 2,4-Dichlorobenzoic acid (DCBA), an allelochemical of cucumber, was almost completely decomposed by plasma treatment. The discharge-treated solution is used as a nutrient solution for cultivating cucumber plants in a rockwool hydroponic system. Plant growth was significantly inhibited by adding DCBA to the nutrient solution, on the other hand, the plants grew healthy using discharge treatment, as shown in Figure 20 [69]. Furthermore, when the feeding gas contains nitrogen, NO2 and NO3 are generated as mentioned already. These ions can act as fertilizers for plants, which improves the growth rate of plants in hydroponic systems, as shown in Figure 21 [107]. Therefore, plasma treatment of nutrient solution for plant cultivation can control the environment of plant growth through the decomposition of toxic organic compounds, the inactivation of pathogenic bacteria, and the nutrient supply, which are promising new applications of plasma. A plasma treatment system for tomato cultivation in a greenhouse was developed, and a pilot test was performed over several months [108]. The CFU of R. solanacearum was kept at a low value, and the infection risk of the plant was significantly reduced during continuous treatment. All plant bodies grow healthily and bear fruits, and the onset of bacterial wilt disease is not observed.

Figure 19.

Photograph of tomato seedlings after 12 days of cultivation in bacteria-contaminated nutrient medium in a hydroponic system.

Figure 20.

Photographs of cucumber seedlings after 18 days and diurnal changes in plant height for seedling cultivated in DCBA-contaminated nutrient medium in a hydroponics system.

Figure 21.

Photograph of Japanese mustard spinach seedlings after 28 days of cultivation using plasma-treated nutrient medium-supplied soil.

Not only radicals such as O3 and OH, but also nitrogen species such as ONOO and O2NOOH can be key inactivation agents that cause cell damage and inactivation under acidic conditions. Because these nitrogen species have a long lifetime in the liquid phase, nitrogen species generated by plasma treatment remain for a long time after nonthermal plasma irradiation. It has been reported that nonthermal plasma-activated water (PAW), water irradiated by nonthermal plasma, can inactivate bacteria on the surface of fruits and vegetables [109] such as strawberries, [110] mushrooms [111] and endives, [112] which contributes to maintaining the freshness of fruits and vegetables.

Advertisement

6. Conclusion

The fundamental physical and chemical properties of plasma discharge generated over a water surface and its applications for the decomposition of persistent compounds are described. Plasma can directly generate powerful oxidizing agents such as hydroxyl radicals, which can realize high-speed decomposition of persistent compounds in wastewater, such as 1,4-dioxane and dichloromethane, without pretreatment of samples. Because of these advantages over other advanced oxidation processes, a pulsed discharge plasma in water and in contact with water has attracted much attention as a promising technology in not only wastewater treatment but also various new application fields such as agriculture.

Authors are grateful to Prof. Y. Minanitani in Yamagata University, Japan for providing a photograph of discharge used in Figure 2(c). This research was supported by a grant-in-aid from the Ministry of Agriculture, Forestry and Fisheries of Japan (A scheme to revitalize agriculture and fisheries in disaster areas through deploying highly advanced technology), by a grants-in-aid for Scientific Research (S), (A), (B), (C) and early-career scientists, from the Japan Society for the Promotion of Science, Grant Numbers 19H05611, 15H02231, 20,370,130, 21 K03994, 18 K13735 and respectively, and by Joint Usage/Research by the Institute of Pulsed Power Science, Kumamoto University.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Attri P, Sarinont T, Kim M, Amano T, Koga K, Cho AE, et al. Influence of ionic liquid and ionic salt on protein against the reactive species generated using dielectric barrier discharge plasma. Scientific Reports. 2015;5:17781. DOI: 10.1038/srep17781
  2. 2. Nakajima A, Uchida G, Kawasaki T, Koga K, Sarinont T, Amano T, et al. Effects of gas flow on oxidation reaction in liquid induced by He/O2 plasma-jet irradiation. Journal of Applied Physics. 2015;118:043301. DOI: 10.1063/1.4927217
  3. 3. Graves DB. The emerging role of reactive oxygen and nitrogen species in redox biology and some implications for plasma applications to medicine and biology. Journal of Physics D: Applied Physics. 2012;45:263001. DOI: 10.1088/0022-3727/45/26/263001
  4. 4. Huang Y, Kou Y, Zheng C, Xu Y, Liu Z, Yan K. Escherichia coli inactivation in water using pulsed discharge. IEEE Transactions on Plasma Science. 2016;44:2654-2655. DOI: 10.1109/TPS.2016.2559802
  5. 5. Kirkpatrick MJ, Dodet B, Odic E. Atmospheric pressure humid argon DBD plasma for the application of sterilization- measurement and simulation of hydrogen, oxygen, and hydrogen peroxide formation. International Journal of Plasma Environmental Science and Technology. 2007;1:96-101
  6. 6. Miklos DB, Remy C, Jekel M, Linden KG, Hübner U. Evaluation of advanced oxidation processes for water and wastewater treatment – A critical review. Water Research. 2018;139:118-131. DOI: 10.1016/j.watres.2018.03.042
  7. 7. Takeuchi N, Yasuoka K. Review of plasma-based water treatment technologies for the decomposition of persistent organic compounds. Japanese Journal of Applied Physics. 2021;60:SA0801. DOI: 10.35848/1347-4065/abb75d
  8. 8. Qin L, Kim K, Takahashi K, Kang J, Takeuchi N, Li OL. N2/Ar plasma-induced surface sulfonation on graphene nanoplatelets for catalytic hydrolysis of cellulose to glucose. Applied Surface Science. 2021;545:149051. DOI: 10.1016/j.apsusc.2021.149051
  9. 9. Takahashi K, Saito Y, Oikawa R, Okumura T, Takaki K, Fujio T. Development of automatically controlled corona plasma system for inactivation of pathogen in hydroponic cultivation medium of tomato. Journal of Electrostatics. 2018;91:61-69. DOI: 10.1016/j.elstat.2017.12.006
  10. 10. Takaki K, Hayashi H, Wang D, Ohshima T. High-voltage technologies for agriculture and food processing. Journal of Physics D: Applied Physics. 2019;52:473001. DOI: 10.1088/1361-6463/ab2e2d
  11. 11. Tanaka H, Ishikawa K. State of the art in medical applications using non-thermal atmospheric pressure plasma. Reviews of Modern Plasma Physics. 2017;1:1-89. DOI: 10.1007/s41614-017-0004-3
  12. 12. Fridman A, Chrokov A, Gutsol A. Non-thermal atmospheric pressure. Journal of Physics D: Applied Physics. 2005;38:R1-R24. DOI: 10.1088/0022-3727/38/2/R01
  13. 13. Huiskamp T. Nanosecond pulsed streamer discharges Part I : Generation , source-plasma interaction and energy-ef fi ciency optimization. Plasma Sources Science and Technology. 2020;29:023002. DOI: 10.1088/1361-6595/ab53c5
  14. 14. Wang D, Namihira T. Nanosecond pulsed streamer discharges : II . Physics , discharge characterization and plasma processing. Plasma Sources Science and Technology. 2020;29:023001. DOI: 10.1088/1361-6595/ab5bf6
  15. 15. Mankowski J, Kristiansen M, Fellow L. A review of short pulse generator technology. IEEE Transactions on Plasma Science. 2000;28:102-108. DOI: 10.1109/27.842875
  16. 16. Wetz DA, Mankowski JJ, Dickens JC, KristiansenM. The impact of field enhancements and charge injetion on the puled breakdown strength of water. IEEE Transactions on Plasma Science. 2006;34:1670-1679. DOI: 10.1109/TPS.2006.881891
  17. 17. Gavrilov IM, Kukhta VR, Lopatin VV, Petrov PG. Dynamics of prebreakdown phenomena in a uniform field. IEEE Transactions on Dielectrics and Electrical Insulation. 1994;1:496-502. DOI: 10.1109/94.300293
  18. 18. Clements JS, Sato M, Davis RH. Preliminary investigation of prebreakdown phenomena and chemical reactions using a pulsed high-voltage discharge in water. IEEE Transactions on Industry Applications. 1987;IA-23:224-235. DOI: 10.1109/TIA.1987.4504897
  19. 19. Bruggeman P, Leys C. Non-thermal plasmas in and in contact with liquids. Journal of Physics D: Applied Physics. 2009;42:053001. DOI: 10.1088/0022-3727/42/5/053001
  20. 20. Takahashi K, Sasaki Y, Mukaigawa S, Takaki K, Fujiwara T, Satta N. Purification of high-conductivity water using gas–liquid phase discharge reactor. IEEE Transactions on Plasma Science. 2010;38:2694-2700. DOI: 0.1109/TPS.2010.2049505
  21. 21. Levko D, Sharma A, Raja LL. Plasmas generated in bubbles immersed in liquids: direct current streamers versus microwave plasma. Journal of Physics D: Applied Physics. 2016;49:285205. DOI: 10.1088/0022-3727/49/28/285205
  22. 22. Hamdan A, Cha MS. The effects of gaseous bubble composition and gap distance on the characteristics of nanosecond discharges in distilled water. Journal of Physics D: Applied Physics. 2016;49:245203. DOI: 10.1088/0022-3727/49/24/245203
  23. 23. Bruggeman PJ, Kushner MJ, Locke BR, Gardeniers JGE, Graham WG, Graves DB, et al. Plasma–liquid interactions: a review and roadmap. Plasma Sources Science and Technology. 2016;25:53002. DOI: 10.1088/0963-0252/25/5/053002
  24. 24. Malik MA. Water purification by plasma which reactors are most energy efficient? Plasma Chemistry and Plasma Processing. 2010;30:21-31. DOI: 10.1007/s11090-009-9202-2
  25. 25. Kanazawa S, Kawano H, Watanabe S, Furuki T, Akamine S, Ichiki R, et al. Observation of OH radicals produced by pulsed discharges on the surface of a liquid. Plasma Sources Science and Technology. 2011;20:034010. DOI: 10.1088/0963-0252/20/3/034010
  26. 26. Hoffer P, Sugiyama Y, Hosseini SHR, Akiyama H, Lukes P, Akiyama M. Characteristics of meter-scale surface electrical discharge propagating along water surface at atmospheric pressure. Journal of Physics D: Applied Physics. 2016;49:415202. DOI: 10.1088/0022-3727/49/41/415202
  27. 27. Xiong Y, Zhang Q, Wandell R, Bresch S, Wang H, Locke BR. Synergistic 1 , 4-dioxane removal by non-thermal plasma followed by biodegradation Synergistic 1,4-dioxane removal by non-thermal plasma followed by biodegradation. Chemical Engineering Journal. 2018;361:519-527. DOI: 10.1016/j.cej.2018.12.094
  28. 28. Wandell RJ, Wang H, Bulusu RKM, Gallan RO, Locke BR. Formation of nitrogen oxides by nanosecond pulsed plasma discharges in gas – Liquid reactors. Plasma Chemistry and Plasma Processing. 2019;39:643-666. DOI: 10.1007/s11090-019-09981-w
  29. 29. Takahashi K, Konno R, Satta N, Takaki K. Estimation of hydroxyl radicals produced by pulsed discharge inside bubble in water using indigo carmine as chemical probe. IEEE Transactions on Plasma Science. 2018;46:663-669. DOI: 10.1109/TPS.2018.2792480
  30. 30. Li J, Sato M, Ohshima T. Degradation of phenol in water using a gas–liquid phase pulsed discharge plasma reactor. Thin Solid Films. 2007;515:4283-4288. DOI: 10.1016/j.tsf.2006.02.070
  31. 31. Takahashi K, Yagi I, Takaki K, Satta N. Development of pulsed discharge inside bubble in water. IEEE Transactions on Plasma Science. 2011;39:2654-2655. DOI: 10.1109/TPS.2011.2164095
  32. 32. Takahashi K, Takayama H, Kobayashi S, Takeda M, Nagata Y, Karashima K, et al. Observation of the development of pulsed discharge inside a bubble under water using ICCD camera. Vacuum. 2020;182:109690. DOI: 10.1016/j.vacuum.2020.109690
  33. 33. Sharma A, Levko D, Raja LL, Cha MS. Kinetics and dynamics of nanosecond streamer discharge in atmospheric-pressure gas bubble suspended in distilled water under saturated vapor pressure conditions. Journal of Physics D: Applied Physics. 2016;49:395205. DOI: 10.1088/0022-3727/49/39/395205
  34. 34. Tachibana K, Takekata Y, Mizumoto Y, Motomura H, Jinno M. Analysis of a pulsed discharge within single bubbles in water under synchronized conditions. Plasma Sources Science and Technology. 2011;20:34005. DOI: 10.1088/0963-0252/20/3/034005
  35. 35. Physics A, Sobota A, Universiteit T, Technische V, Eindhoven U, Philips M. Speed of streamers in argon over a flat surface of a dielectric. Journal of Physics D: Applied Physics. 2008;42:015211. DOI: 10.1088/0022-3727/42/1/015211
  36. 36. Babaeva NY, Tereshonok DV, Naidis GV. Fluid and hybrid modeling of nanosecond surface discharges: effect of polarity and secondary electrons emission. Plasma Sources Science and Technology. 2016;25:044008. DOI: 10.1088/0963-0252/25/4/044008
  37. 37. Bruggeman P, Guns P, Degroote J, Vierendeels J. Influence of the water surface on the glow-to-spark transition in a metal-pin-to-water electrode system. Plasma Sources Science and Technology. 2008;17:045014. DOI: 10.1088/0963-0252/17/4/045014
  38. 38. Takahashi K, Takeda M, Konno R, Takaki K, Satta N. Influence of electric parameters on hydroxyl radical production by positive pulsed discharge inside of a bubble in water. IEEE Transactions on Plasma Science. 2019;47:1105-1113. DOI: 10.1109/TPS.2018.2883767
  39. 39. Takahashi K, Mukaigawa S, Takaki K, Fujiwara T, Satta N. Water purification using non-thermal plasma driven by blumlein-line stacked pulsed power generator. Journal of Plasma and Fusion Research SERIES. 2009;8:1459-1462
  40. 40. Minamitani Y, Yamada T. Investigation of the influence of droplets to streamer discharge in water treatment by pulsed discharge in air spraying. IEEE Transactions on Plasma Science. 2016;44:2173-2180. DOI: 10.1109/TPS.2016.2592515
  41. 41. Stefanovic I, Bibinov N, Deryugin A, Vinoradov I, Napartovich A, Wiesemann K. Kinetics of ozone and nitric oxides in dielectric barrier discharges in O2/NOx and N2/O2/NOx mixtures. Plasma Sources Science and Technology. 2001;10:406-416. DOI: 10.1088/0963-0252/10/3/303
  42. 42. Luo Y, Lietz AM, Yatom S, Kushner MJ, Bruggeman PJ. Plasma kinetics in a nanosecond pulsed filamentary discharge sustained in Ar – H2O and H2O. Journal of Physics D: Applied Physics. 2019;52:044003. DOI: 10.1088/1361-6463/aaeb14
  43. 43. Liu Y, Tan Z, Chen X, Li X, Wang X. A numerical investigation on the effects of water vapor on electron energy and OH production in atmospheric-pressure He/H2O and Ar/H2O plasma jets. IEEE Transactions on Plasma Science. 2019;47:1593-1604. DOI: 10.1109/TPS.2019.2896060
  44. 44. Shirafuji T, Murakami T. Contribution of electrons, Ar(3P0,2), H2O+, and H3O+ to production of OH(A2Σ+) in a micro-dielectric barrier discharge of Ar/H2O. Japanese Journal of Applied Physics. 2014;54:01AC03. DOI: 10.7567/JJAP.54.01AC03
  45. 45. Verreycken T, Schram DC, Leys C, Bruggeman P. Spectroscopic study of an atmospheric pressure dc glow discharge with a water electrode in atomic and molecular gases. Plasma Sources Science and Technology. 2010;19:045004. DOI: 10.1088/0963-0252/19/4/045004
  46. 46. Komuro A, Ono R, Oda T. Behaviour of OH radicals in an atmospheric-pressure streamer discharge studied by two-dimensional numerical simulation. Journal of Physics D: Applied Physics. 2013;46:175206. DOI: 10.1088/0022-3727/46/17/175206
  47. 47. Bruggeman P, Liu J, Degroote J, Kong MG. DC excited glow discharges in atmospheric pressure air in pin-to-water electrode. Journal of Physics D: Applied Physics. 2008;41:215201. DOI: 10.1088/0022-3727/41/21/215201
  48. 48. Wakisaka S, Tsuda K, Takahashi K, Satoh K. Mechanism of pH variation and H 2 O 2 generation in water exposed to pulsed discharge plasma. IEEE Transactions on Plasma Science. 2019;47:1083-1088. DOI: 10.1109/TPS.2018.2866282
  49. 49. Atkinson R, Baulch DL, Cox RA, Hampson RF, Kerr JA, Troe J. Evaluated kinetic and photochemical data for atmospheric chemistry: Supplement III. IUPAC subcommittee on gas kinetic data evaluation for atmospheric chemistry. Journal of Physical and Chemical Reference Data. 1989;18:881. DOI: 10.1063/1.555832
  50. 50. Gerasimov GY, Shatalov OP. Kinetic mechanism of combustion of hydrogen–oxygen mixtures. Journal of Engineering Physics and Thermophysics. 2013;86:987-995. DOI: 10.1007/s10891-013-0919-7
  51. 51. Yagi I, Ono R, Oda T, Takaki K. Two-dimensional LIF measurements of humidity and OH density resulting from evaporated water from a wet surface in plasma for medical use. Plasma Sources Science and Technology. 2014;24:015002. DOI: 10.1088/0963-0252/24/1/015002
  52. 52. Takeuchi N, Ando M, Yasuoka K. Investigation of the loss mechanisms of hydroxyl radicals in the decomposition of organic compounds using plasma generated over water. Japanese Journal of Applied Physics. 2015;54:116201. DOI: 10.7567/JJAP.54.116201
  53. 53. Chen C, Liu DX, Liu ZC, Yang AJ, Chen HL, Shama G, et al. A model of plasma-biofilm and plasma-tissue interactions at ambient pressure. Plasma Chemistry and Plasma Processing. 2014;34:403-441. DOI: 10.1007/s11090-014-9545-1
  54. 54. Rumbach P, Bartels DM, Go DB. The penetration and concentration of solvated electrons and hydroxyl radicals at a plasma-liquid interface. Plasma Sources Science and Technology. 2018;27:115013. DOI: 10.1088/1361-6595/aaed07
  55. 55. Kanazawa S, Kawano H, Watanabe S, Furuki T, Akamine S, Ichiki R, et al. Observation of OH radicals produced by pulsed discharges on the surface of a liquid. Plasma Sources Science and Technology. 2011;20:34010. DOI: 10.1088/0963-0252/20/3/034010
  56. 56. Gomes A, Fernandes E, Lima JLFC. Fluorescence probes used for detection of reactive oxygen species. Journal of Biochemical and Biophysical Methods. 2005;65:45-80. DOI: 10.1016/j.jbbm.2005.10.003
  57. 57. Shiraki D, Ishibashi N, Takeuchi N. Quantitative estimation of OH radicals reacting in liquid using a chemical probe for plasma in contact with liquid. IEEE Transactions on Plasma Science. 2016;44:3158-3163. DOI: 10.1109/TPS.2016.2603157
  58. 58. Knipping EM, Dabdub D. Modeling Cl2 formation from aqueous NaCl particles: Evidence for interfacial reactions and importance of Cl2 decomposition in alkaline solution. Journal of Geophysical Research. 2002;107:4360. DOI: 10.1029/2001JD000867
  59. 59. Yuchen H, Uehara S, Tanaka H, Nishiyama H. Numerical modelling and simulation of chemical reactions in a nano-pulse discharged bubble for water treatment. Plasma Science and Technology. 2016;18:924-934. DOI: 10.1088/1009-0630/18/9/09
  60. 60. Ono R, Oda T. Dynamics of ozone and OH radicals generated by pulsed corona discharge in humid-air flow reactor measured by laser spectroscopy. Journal of Applied Physics. 2003;93:5876-5882. DOI: 10.1063/1.1567796
  61. 61. Nakagawa Y, Ono R, Oda T. Effect of discharge polarity on OH density and temperature in coaxial-cylinder barrier discharge under atmospheric pressure humid air. Japanese Journal of Applied Physics. 2018;57:096103. DOI: 10.7567/JJAP.57.096103
  62. 62. Staehelln J, Holgne J. Decomposition of ozone in water: Rate of initiation by hydroxide ions and hydrogen peroxide. Environmental Science & Technology. 1982;16:676-681. DOI: 10.1021/es00104a009
  63. 63. Kim HH, Prieto G, Takashima K, Katsura S, Mizuno A. Performance evaluation of discharge plasma process for gaseous pollutant removal. Journal of Electrostatics. 2002;55:25-41. DOI: 10.1016/S0304-3886(01)00182-6
  64. 64. Soloshenko IA, Tsiolko VV, Pogulay SS, Kalyuzhnaya AG, Bazhenov VY, Shchedrin AI. Effect of water adding on kinetics of barrier discharge in air. Plasma Sources Science and Technology. 2009;18:045019. DOI: 10.1088/0963-0252/18/4/045019
  65. 65. Machala Z, Tarabova B, Hensel K, Spetolikova E, Sikurova L, Lukes P. Formation of ROS and RNS in water electro-sprayed through transient spark discharge in air and their bactericidal effects. Plasma Processes and Polymers. 2013;10:649-659. DOI: 10.1002/ppap.201200113
  66. 66. Régimbal JM, Mozurkewich M. Peroxynitric acid decay mechanisms and kinetics at low pH. The Journal of Physical Chemistry. A. 1997;101:8822-8829. DOI: 10.1021/jp971908n
  67. 67. Warneck P. Chemical reactions in clouds. Fresenius' Journal of Analytical Chemistry. 1991;340:585. DOI: 10.1007/BF00322434
  68. 68. Hoigne J, Bader H, Haag WR, Staehelin J. Rate constants of reactions of ozone with organic and inorganic compounds in water—III. Inorganic compounds and radicals. Water Research. 1985;19:993-1004. DOI: 10.1016/0043-1354(85)90368-9
  69. 69. Takahashi K, Kawamura S, Takada R, Takaki K, Satta N, Fujio T. Influence of a plasma-treated nutrient solution containing 2,4-dichlorobenzoic acid on the growth of cucumber in a hydroponic system. Journal of Applied Physics. 2021;129:143301. DOI: 10.1063/5.0044328
  70. 70. Hug SJ, Leupin O. Iron-catalyzed oxidation of Arsenic(III) by oxygen and by hydrogen perxide: pH-dependent formation of oxidants in the fenton reaction. Environmental Science & Technology. 2003;37:2734-2742. DOI: 10.1021/es026208x
  71. 71. Kremer ML. Mechanism of the Fenton reaction. Evidence for a new intermediate. Physical Chemistry Chemical Physics. 1999;1:3595-3605. DOI: 10.1039/A903915E
  72. 72. White EM, Vaughan PP, Zepp RG. Role of the photo-Fenton reaction in the production of hydroxyl radicals and photobleaching of colored dissolved organic matter in a coastal river of the southeastern United States. Aquatic Sciences. 2003;65:402-412. DOI: 10.1007/s00027-003-0675-4
  73. 73. Takahashi K, Takaki K, Satta N. Water remediation using pulsed power discharge under water with advanced oxidation process. Journal of Advanced Oxidation Technologies. 2012;15:365-373. DOI: 10.1515/jaots-2012-0216
  74. 74. Liu Q, Schurter LM, Muller CE, Aloisio S, Francisco JS, Margerum DW. Kinetics and mechanisms of aqueous ozone reactions with bromide, sulfite, hydrogen sulfite, iodide, and nitrite ions. Inorganic Chemistry. 2001;40:4436-4442. DOI: 10.1021/ic000919j
  75. 75. Buxton GV, Greenstock CL, Helman WP, Ross AB. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals in aqueous solution. Journal of Physical and Chemical Reference Data. 1987;17:513. DOI: 10.1063/1.555805
  76. 76. Ross F, Ross AB. Selected specific rates of reactions of transients from water in aqueous solution. In: III. Hydroxyl Radical and Perhydroxyl Radical and Their Radical Ions. United States: N: Web; 1977. DOI: 10.2172/6635906
  77. 77. Logan SR. Redox reactions of organic radicals with ferrocene/ferricenium species in aqueous solution. Part 1. Radicals derived from carboxylic acids. Journal of the Chemical Society, Perkin Transactions. 1989;2:751-754. DOI: 10.1139/v91-081
  78. 78. Hoigne J, Bader H. Rate constants of reactions of ozone with organic and inorganic compounds in waterII Dissociating organic compounds. Water Research. 1983;17:158-194. DOI: 10.1016/0043-1354(83)90098-2
  79. 79. Neta P, Huie RE, Ross AB. Rate constants for reactions of inorganic radicals in aqueous solution rate · constants for reactions of inorganic radicals in aqueous solution. Journal of Physical and Chemical Reference Data. 2009;17:1027-1284. DOI: 10.1063/1.555808
  80. 80. Abe A. Distribution of 1, 4-dioxane in relation to possible sources in the water environment. Science of the Total Environment. 1999;227:41-47. DOI: 10.1016/S0048-9697(99)00003-0
  81. 81. Zenker MJ, Borden RC, Barlaz MA. Occurrence and treatment of 1,4-dioxane in aqueous environments. Environmental Engineering Science. 2003;20:423-443. DOI: 10.1089/109287503768335913
  82. 82. Khan FI, Ghoshal AK. Removal of volatile organic compounds from polluted air. Journal of Loss Prevention in the Process Industries. 2000;13:527-545. DOI: 10.1016/S0950-4230(00)00007-3
  83. 83. Huang B, Lei C, Wei C, Zeng G. Chlorinated volatile organic compounds (Cl-VOCs) in environment — sources, potential human health impacts, and current remediation technologies. Environment International. 2014;71:118-138. DOI: 10.1016/j.envint.2014.06.013
  84. 84. Bian W, Song X, Liu D, Zhang J, Chen X. Actions of nitrogen plasma in the 4-chrolophenol degradation by pulsed high-voltage discharge with bubbling gas. Chemical Engineering Journal. 2013;219:385-394. DOI: 10.1016/j.cej.2012.12.074
  85. 85. Nunnally T, Tsangaris A, Rabinovich A, Nirenberg G. Gliding arc plasma oxidative steam reforming of a simulated syngas containing naphthalene and toluene. International Journal of Hydrogen Energy. 2014;39:11976-11989. DOI: 10.1016/j.ijhydene.2014.06.005
  86. 86. Yang J, Zeng D, Hassan M, Ma Z, Dong L, Xie Y, et al. Efficient degradation of Bisphenol A by dielectric barrier discharge non-thermal plasma: Performance, degradation pathways and mechanistic consideration. Chemosphere. 2021;286:131627. DOI: 10.1016/j.chemosphere.2021.131627
  87. 87. Lei W, Huifen Z, Xin Y. Dechlorination and decomposition of trichloroacetic acid by glow discharge plasma in aqueous solution. Electrochimica Acta. 2014;115:332-336. DOI: 10.1016/j.electacta.2013.10.160
  88. 88. Ghodbane H, Hamdaoui O, Vandamme J, Durme JV, Vanraes P, Leys C, et al. Degradation of AB25 dye in liquid medium by atmospheric pressure non-thermal plasma and plasma combination with photocatalyst TiO2. Open Chemistry. 2014;13:325-331. DOI: 10.1515/chem-2015-0040
  89. 89. Shirafuji T, Ishida Y, Nomura A, Hayashi Y, Goto M. Reaction mechanisms of methylene-blue degradation in three-dimensionally integrated micro-solution plasma. Jpn. Journal of Applied Physics. 2017;56:06HF02. DOI: 10.7567/JJAP.56.06HF02
  90. 90. Iwabuchi M, Takahashi K, Takaki K, Satta N. Influence of sodium carbonate on decomposition of formic acid by pulsed discharge plasma inside bubble in water. Japanese Journal of Applied Physics. 2016;55:07LF02. DOI: 10.7567/JJAP.55.07LF02
  91. 91. Takahashi K, Takayama H, Yagi I, Takaki K, Satta N. Decomposition process of volatile organic compounds dissolved into water by pulsed discharge inside bubble. Japanese Journal of Applied Physics. 2020;59:SHHA06. DOI: 10.35848/1347-4065/ab72cd
  92. 92. Reife A, Freeman HS. Environmental Chemistry of Dyes and Pigments. New York, NY: John Wiley & Sons; 1996. pp. 50-59
  93. 93. Backett MA, Hua I. Elucidation of the 1,4-dioxane decomposition pathway at discrete ultrasonic frequencies. Environmental Science & Technology. 2000;34:3944-3953. DOI: doi.org/10.1021/es000928r
  94. 94. Stefan MI, Bolton JR. Mechanism of the degradation of 1,4-dioxane in dilute aqueous solution using the UV/hydrogen peroxide process. Environmental Science & Technology. 1998;32:1588-1595. DOI: 10.1021/es970633m
  95. 95. Höbel B, Sonntag C. OH-Radical induced degradation of ethylenediaminetetraacetic acid(EDTA) in aqueous solution; a pulse raiolysis study. Journal of the Chemical Society, Perkin Transactions. 1998;2:509-513. DOI: 10.1039/A708167G
  96. 96. Thomas JK. Rates of reaction of the hydroxyl radical. Transactions of the Faraday Society. 1966;61:702. DOI: 10.1039/TF9656100702
  97. 97. Haag WR, Yao CCCD. Rate constants for reaction off hydroxyl radicals with several drinking water contaminants. Environmental Science & Technology. 1992;26:1005-1013. DOI: 10.1021/es00029a021
  98. 98. Fitzsimmons C, Ismail F, Whitehead JC, Wilman JJ. The chemistry of dichloromethane destruction in atmospheric-pressure gas streams by a dielectric packed-bed plasma reactor. The Journal of Physical Chemistry. A. 2000;104:6032-6038. DOI: doi.org/10.1021/jp000354c
  99. 99. Kundu SK, Kennedy EM, Mackie JC, Holdsworth CI, Molloy TS, Gaikwad VV, et al. Study on the reaction of CCl2F2 with CH4 in a dielectric barrier discharge nonequilibrium plasma. Plasma Processes and Polymers. 2013;10:912-921. DOI: 10.1002/ppap.201300053
  100. 100. Gaikwad V, Kennedy E, Mackie J, Holdsworth C, Molloy S, Kundu S, et al. Reaction of dichloromethane under non-oxidative conditions in a dielectric barrier discharge reactor and characterisation of the resultant polymer. Chemical Engineering Journal. 2016;290:499-506. DOI: 10.1016/j.cej.2015.12.105
  101. 101. Orlandini I, Riedel U. Chemical kinetics of NO removal by pulsed corona discharges. Journal of Physics D: Applied Physics. 2000;33:2467
  102. 102. Takaki K, Takahashi K, Hamanaka D, Yoshida R, Uchino T. Function of plasma and electrostatics for keeping quality of agricultural produce in post-harvest stage. Japanese Journal of Applied Physics. 2021;60:010501. DOI: 10.35848/1347-4065/abcc13
  103. 103. Takaki K, Hayashi N, Wang D, Ohshima T. High-voltage technologies for agriculture and food processing. Journal of Physics D: Applied Physics. 2019;52:473001. DOI: 10.1088/1361-6463/ab2e2d
  104. 104. Puac N, Gherardi M, Shiratani M. Plasma agriculture : A rapidly emerging field. Plasma Processes and Polymers. 2018;15:1700174. DOI: 10.1002/ppap.201700174
  105. 105. Ranieri P, Kizer J, Gatiboni L, Hernández R, Pierce MR, Stapelmann K, et al. Plasma agriculture : Review from the perspective of the plant and its ecosystem. Plasma Processes and Polymers. 2020;18:e2000162. DOI: 10.1002/ppap.202000162
  106. 106. Okumura T, Saito Y, Takano K, Takahashi K, Takaki K, Satta N, et al. Inactivation of bacteria using discharge plasma under liquid fertilizer in a hydroponic culture system. Plasma Medicine. 2017;6:247-254. DOI: 10.1615/PlasmaMed.2016018683
  107. 107. Takahata J, Takaki K, Satta N, Takahashi K. Improvement of growth rate of plants by bubble discharge in water. Japanese Journal of Applied Physics. 2015;54:01AG07. DOI: 10.7567/JJAP.54.01AG07
  108. 108. Takahashi K, Saito Y, Oikawa R, Okumura T, Takaki K, Fujio T. Development of automatically controlled corona plasma system for inactivation of pathogen in hydroponic cultivation medium of tomato. Journal of Electrostatics. 2018;91:61-69. DOI: 10.1016/j.elstat.2017.12.006
  109. 109. Schnabel U, Andrasch M, Weltmann K. Antimicrobial efficiency of non-thermal atmospheric pressure plasma processed water (PPW) against agricultural relevant bacteria suspensions. International Journal of Environmental & Agriculture Research. 2016;2:212-224
  110. 110. Ma R, Wang G, Tian Y, Wang K, Zhang J, Fang J. Non-thermal plasma-activated water inactivation of food-borne pathogen on fresh produce. Journal of Hazardous Materials. 2015;300:643-651. DOI: 10.1016/j.jhazmat.2015.07.061
  111. 111. Xu Y, Tian Y, Ma R, Liu Q, Zhang J. Effect of plasma activated water on the postharvest quality of button mushrooms, Agaricus bisporus. Food Chemistry. 2016;197:436-444. DOI: 10.1016/j.foodchem.2015.10.144
  112. 112. Andrasch M, Stachowiak J, Schlüter O, Schnabel U, Ehlbeck J. Scale-up to pilot plant dimensions of plasma processed water generation for fresh-cut lettuce treatment. Food Packaging and Shelf Life. 2017;14:40-45. DOI: 10.1016/j.fpsl.2017.08.007

Written By

Katsuyuki Takahashi, Koichi Takaki and Naoya Satta

Submitted: 31 July 2021 Reviewed: 04 November 2021 Published: 10 December 2021