Next Article in Journal
Steam Reforming of Bio-Compounds with Auto-Reduced Nickel Catalyst
Previous Article in Journal
Improving the Stability of Cold-Adapted Enzymes by Immobilization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Abatement of VOCs Using Packed Bed Non-Thermal Plasma Reactors: A Review

by
Savita K. P. Veerapandian
*,
Christophe Leys
,
Nathalie De Geyter
and
Rino Morent
Department of Applied Physics, Research Unit Plasma Technology, Faculty of Engineering and Architecture, Ghent University, Sint-Pietersnieuwstraat 41, 9000 Ghent, Belgium
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(4), 113; https://doi.org/10.3390/catal7040113
Submission received: 20 December 2016 / Revised: 6 April 2017 / Accepted: 7 April 2017 / Published: 12 April 2017

Abstract

:
Non thermal plasma (NTP) reactors packed with non-catalytic or catalytic packing material have been widely used for the abatement of volatile organic compounds such as toluene, benzene, etc. Packed bed reactors are single stage reactors where the packing material is placed directly in the plasma discharge region. The presence of packing material can alter the physical (such as discharge characteristics, power consumption, etc.) and chemical characteristics (oxidation and destruction pathway, formation of by-products, etc.) of the reactor. Thus, packed bed reactors can overcome the disadvantages of NTP reactors for abatement of volatile organic compounds (VOCs) such as lower energy efficiency and formation of unwanted toxic by-products. This paper aims at reviewing the effect of different packing materials on the abatement of different aliphatic, aromatic and chlorinated volatile organic compounds.

1. Introduction

Abatement of low concentration of volatile organic compounds (VOCs) from air is of great interest in the past two decades for environmental remediation. VOCs are organic compounds with very high vapor pressure, which are found in trace amounts in the atmosphere (typically less than 1000 ppm). Emission of VOCs into the atmosphere is of great concern due to its adverse effect on humans (carcinogen, affects central nervous system (CNS), causes respiratory problem, etc.) and the environment (ozone depletion, photochemical smog, global warming, etc.). The sources and the adverse effects of different VOCs on human and environment are shown in Table 1.
Conventional techniques like adsorption [1], absorption [2,3], condensation [4,5], catalytic oxidation [6,7], photocatalysis [8,9,10], membrane purification [11] and biological treatment [12,13] which have been used for removal of VOCs have disadvantages such as low efficiency and high energy consumption. The merits and demerits of various technologies available for the abatement of VOCs are reviewed in the work by khan et al. [14], Parmar et al. [15] and Luengas et al. [16]. For example, in catalytic oxidation, the catalyst has to be heated and maintained at higher temperature throughout the process [17]. In thermal oxidation, large amount of energy is used to heat the large volume of gas containing very low concentration of VOCs [18]. Thus, the above mentioned techniques are not cost effective to treat the waste gas when the concentration of VOC is less than 1000 ppm, as it is difficult to maintain the adiabatic conditions [19]. Alternatively, non-thermal plasma (NTP) technology is cost effective in converting diluted VOCs in large volume of air into less toxic substances [20,21,22,23]. NTP technology utilizes the supplied energy to create high energy electrons whose temperature is in the range of 10,000–250,000 K (1–25 eV) [24]; whereas the temperature of other particles remains close to room temperature [25,26]. Thus, the electrons are not in thermal equilibrium with other particles and the overall temperature of the system is maintained close to room temperature. These energetic electrons collide with carrier gas molecules, which is predominantly air and generate highly reactive species such as ground and excited atomic oxygen and nitrogen, and vibrationally and electronically excited oxygen and nitrogen molecules. These reactive species in turn oxidize the diluted VOCs into less toxic substances such as H2O, CO2 and other by-products.
Although different types of NTP discharge at atmospheric pressure and room temperature such as corona discharge [27,28,29,30], surface discharge (SD) [31], microwave discharge [32], dielectric barrier discharge (DBD) [33,34,35,36,37], packed bed DBD (PBDBD) [38,39,40] have been widely investigated as an alternative technology to remove VOCs from exhaust gas, formation of unwanted by-products and high energy consumption are the main bottlenecks for the commercialization of the technology. For the development of better NTP technology for abatement of VOCs, the two important factors to be considered in addition to decomposition efficiency are (i) reducing energy consumption and (ii) controlling the formation of unwanted and toxic by-products such as newly formed other VOCs, ozone, NOx, SOx [41]. It has been widely reported in the literature that DBD reactors packed with non-catalytic or catalytic materials enhance the energy efficiency [20,42] and reduce the formation of unwanted by-products by total oxidation of aliphatic (methane, butane, ethylene, formaldehyde, acetaldehyde, acetone, bromomethane), aromatic (benzene, toluene, styrene, xylene, chlorobenzene) and halogenated (carbon tetrachloride, dichloromethane, perfluoroethane) VOCs. The effect of different packing materials on the discharge characteristics of the reactor, decomposition efficiency of different VOCs and formation of by-products are summarized in Table 2.
The presence of packing material in the discharge region of the reactor enhances the electric field near the contact points between the dielectric pellets and increases the residence time of VOCs in the discharge region resulting in enhanced removal efficiency [43] and deep oxidation [44] of VOCs respectively. The total oxidation of VOCs in packed bed reactors depends on the properties of packing materials (dielectric constant, size, shape, surface properties and catalytic activity), presence of humidity, energy density and nature of VOC itself. For example, aromatic compound with aside chain requires lower energy for complete oxidation when compared to the aromatic compound without sidechain like benzene [45]. This is because the ionization potential of benzene is higher than aromatic compounds with sidechain like xylene or toluene [46].
In this review, the overview of the literature on the packed bed dielectric barrier discharge (PBDBD) reactors for the abatement of different volatile organic compounds is presented. In the first part, different types of packed bed reactors which have been used for VOC removal are reviewed. In the second part, the influence of different properties of the packing materials such as dielectric constant, packing material shape and size, surface properties (porosity and surface area) on the performance of packed bed reactors for the decomposition of different VOCs are reported. The effect of addition of catalysts or catalytic packing material on decomposition efficiency (η), carbon balance (the ratio between the carbon converted from VOC to gaseous by-products and the total carbon converted in the process), CO2 selectivity (the ratio between the carbon converted from VOC to CO2 and the total carbon converted in the process) and by-product formation is also reviewed.

2. Packed Bed Reactors

Packed bed reactors are single stage reactors in which a non-catalytic/catalytic packing material is located in the discharge region of the reactor. The schematic of the most commonly used configuration of packed bed reactors are shown in Figure 1. The PBDBD reactors used for the abatement of VOCs consist of a discharge (high voltage) electrode and a ground electrode separated by one or two dielectric layers and the space between the dielectric layer and one of the electrode is filled with the packing material. Quartz or glass tubes are widely used as dielectric barrier in the packed bed reactors used for the decomposition of VOCs.
The high voltage (or discharge) electrode can be made of stainless steel rod [47,48,49,50,51], stainless steel wire [52,53], stainless steel spring [54,55], stainless steel bolt [56,57], aluminum rod [58], copper rod [59], molybdenum rod [60] or tungsten wire [43,61,62,63]. The surface area and geometry of the inner electrode plays an important role in the discharge characteristics of the plasma reactor. For example, increasing the diameter of the inner electrode, reduces the discharge gap and increases the superficial area resulting in a uniform distribution of microdischarges and a higher number of energetic secondary electrons [61] respectively. The geometry of the electrode affects the number of microdischarges per cycle of applied voltage [64]. Using a bolt instead of a wire or rod as high voltage electrode enhances electric field at the sharp edges of the bolt and increases the number of microdischarges resulting in higher removal and power efficiency [37,56,65]. In PBDBD reactors, the length of the ground electrode determines the effective length of the reactor. The ground electrode can be made of aluminum tape [47,52,55], copper mesh [52,58,63,66], stainless steel mesh [43,56,60], iron mesh [61], aluminum mesh [50], silver paste [51,52], stainless steel foil [53], brass wire [62] or aluminum foil [59,67,68]. Using metal wire, tape or gauze as ground electrode can form void between ground electrode and dielectric barrier which results in a corona discharge in the void. The energy consumed by the corona discharge in the void does not contribute to the decomposition of VOC and results in increased energy consumption and decreased conversion efficiency. In literature, it has been mentioned that painting silver paste on the wall of the discharge tube [41,52,69] prevents the formation of voids and improves the conductivity between electrode.
The DBD reactor with packing material operates with a small volume fraction of plasma when compared to the unpacked reactor [70] and this packing material can be non-catalytic or catalytic. The most commonly used packing materials for the abatement of VOCs are (i) non catalytic materials such as glass, Al2O3, TiO2, MnO2, activated carbon, graphene oxide (GO), Raschig ring (RR), Zeolites (HZSM-5, ferrierite, H-Y), glass wool (GW), molecular sieve (MS-3A, MS-4A, MS-5A, MS-13X, OMS-2), ferroelectric materials like BaTiO3, NaNO2, SrTiO3, CaTiO3, Mg2TiO4 and (ii) catalytic materials such as metal oxides and noble metal loaded metal oxides.
In PBDBD reactors, when high voltage is applied across the dielectric barrier, the dielectric packing material is polarized which enhances the electric field around the contact points resulting in a partial discharge [71,72]. As shown in Figure 2, the presence of packing materials in the discharge region maximizes the number of micro streamers and thus producing more high energy electrons [18]. These energetic electrons produce reactive intermediates resulting in increased VOC removal efficiency when compared to the unpacked reactors [57,73,74]. Surface discharges produced along the surface of the packing material may distribute the plasma more evenly and intensively throughout the reactor [75] and the catalytic reactions occur on the surface of the packing material. Thus, the use of packed bed reactor gives the advantage of uniform distribution of gas flow and discharge [76]. In packed bed reactors, absorbents are also used as packing material which increase the residence time and minimize the volume of free gas in the discharge region [77] resulting in increased collisional probability between VOC and active species.
The presence of packing material enhances the energy efficiency of the NTP reactor. The presence of packing material in the discharge region reduces the breakdown voltage required for ignition of plasma by enhancing the electric field near the contact points. As shown in Figure 3, the power consumption of the packed reactors are smaller when compared to the unpacked reactor [78] under the same operational conditions. The presence of the packing material in the reactor can reduce the output power and output current [79] for the same applied voltage due to the higher space charge effect. However the average power density (=input power/gas space) of the packed reactor is higher than the unpacked reactor as the volume of the plasma zone in the packed reactor is smaller than the unpacked one. The performance of the non-packed and packed reactors can also be compared using the energy effectiveness (=g VOC removed/kWh). In [79], Lin et al. showed that the energy effectiveness of a glass beads packed bed reactor is 1.5 times higher when compared to the unpacked reactor. The specific energy density required for the non-packed reactor is quite higher than the packed reactor to obtain similar decomposition efficiency [80].
The different kinds of power source used to provide energy for the packed bed plasma reactor are alternating current (ac) or bipolar pulsed power sources [54]. Although a bipolar pulsed power source is expensive, it has advantages over ac power source such as instantaneous energy input [54] and can avoid the accumulation of net charges on the surface of dielectrics [54]. In [81], Duan et al. compared the traditional ac power supply of 50 Hz and bipolar pulse power supply of 50 Hz (peak-to-peak voltage Vpp = 0 to 30 kV; pulse width 100 ns and pulse rise rate 500 V·(ns)−1). In this work, it is shown that the discharge current produced by bipolar pulsed power supply is higher and longer when compared to ac power supply with same frequency and applied voltage [81] and the bipolar pulsed power can enhance the chlorobenzene removal efficiency by 1.6 times when compared to ac power supply. However, for economic reason, ac power supply is widely used.
The power consumed by packed bed reactors for the discharge can be calculated using Lissajous curve [25] or by integrating the current and voltage waveform over one period [82,83]. In Lissajous method, the voltage applied on the high voltage electrode is measured using a high voltage probe and the dissipated charge is measured across a capacitor of capacitance C m , which is connected in series with the ground electrode. The energy deposited in the plasma reactor per cycle is calculated from the area S of the V-Q Lissajous (Figure 2a,b) diagram which has been recorded using a digital oscilloscope. The average power P during discharge is calculated by multiplying the energy deposited per cycle by ac power frequency f [84],
P = S f
In the second method, the average power, P, of the discharge is calculated using following equation from the measurement of the current and the voltage over one cycle.
P = 1 T t t + T I ( t ) V ( t ) d t
where V ( t ) and I ( t ) are instantaneous voltage and current. The discharge current, I, is measured through a resistor in series with the grounded electrode. The applied voltage V(t), and the current variations I(t) with time are visualized using a digital oscilloscope. The specific input energy (SIE) which is widely used in literature to study the energy efficiency of the process and reactor in abatement of VOC is defined as the energy deposited in the flue gas per unit volume and it is given by [85]:
S I E   ( J / L ) =   P   ( W ) F l o w   r a t e   ( L / s )
The frequency of the ac power supply also influences the energy efficiency of the DBD reactor. Ogata et al. [66,86] reported that the conversion of benzene decreases with increase in ac frequency applied to the DBD reactor. In [65], it has been reported that the energy efficiency for trichloroethylene (TCE) removal is higher with a 50 Hz ac power supply when compared to higher frequencies such as 500 and 2000 Hz. This decrease in energy efficiency at higher frequency is mainly due to the significant convection [87] (thermal transfer between plasma and dielectric surface) and displacement [87] (the current generated by charging the capacitance of the DBD reactor) current which are not useful for the decomposition of the chemical compounds.

3. Properties of Packing Material

3.1. Dielectric Constant

When an electric field is applied across a dielectric material, the factor by which the effective electric field is reduced by polarization of the dielectric material itself is the dielectric constant. The dielectric constant of a packing material determines the amount of energy the reactor can store during a single discharge [88]. When the applied ac voltage is beyond the breakdown voltage of the interstitial space between packing material, a discharge plasma is formed. The electric field in these interstitial space is amplified by the presence of packing material near the contact points [22] of the packing material and the electric field is higher for the pellets of higher dielectric constant due to boundary conditions of electric flux [89]. The applied voltage and discharge current waveforms for BaTiO3 pellets of different dielectric constant (εs = 10,000, 5000 and 660) are shown in Figure 4. The discharge current increases with increase in the dielectric constant of the pellet and it is maximum for the pellet of dielectric constant 104.
Packing material with a higher dielectric constant decreases the breakdown voltage [91] and enhances the electric field resulting in enhanced VOC removal efficiency [92]. Zheng et al. [88] pointed out that the packing material with higher dielectric constant (γ-Al2O3) enhances the removal efficiency of acetone when compared to glass beads. When an external ac voltage beyond onset voltage is applied across a dielectric barrier with higher dielectric constant, an intense electric field is produced in the area near the contact point of the pellets and results in a combination of surface discharges on the packing pellets and filamentary discharges in the voids [43,93]. As shown in Figure 5, ferroelectric packing materials with dielectric constant ε r 1100 are energy efficient in benzene removal (amount of benzene removed per watt) [86]. In [61], Liang et al. reported that the BaTiO3 packed bed reactor has the highest toluene removal efficiency when compared to a NaNO2 PBDBD reactor, as the spontaneous polarization intensity of BaTiO3 is higher than NaNO2. Thus, the usage of ferroelectric pellets enhances the electric field (by a factor of 10–250) by polarization [72], resulting in the formation of higher-energetic species in the discharge zone [94] and thus an enhanced VOC removal efficiency.
Apart from VOC removal efficiency, the dielectric constant of the packing material also plays an important role in the formation of by-products such as ozone, NOx. According to Futamura et al. [41], the nature of the catalyst or the packing material has less effect on the initial conversion of VOC. However, it has more influence on the reaction of intermediates formed leading to a difference in the carbon balance and CO2 selectivity. Ogata et al. [86] reported that the ferroelectric packing material with lower dielectric constant suppresses the formation of NOx, but produces large amount of ozone. Coating dielectric material with metals of higher conductivity reduces the polarization of the pellets and thus the decomposition efficiency. In [71], it is reported that BaTiO3 pellets showed better removal efficiency when compared to 0.2 wt % Pt coated BaTiO3 due to the increased conductivity of the pellets by platinum coating. The dielectric constant of the material varies as a function of temperature. For example, the dielectric constant of the ferroelectric packing material Urashima et al. [92] reported that the temperature of the DBD reactor and thus the packing material increases during the operation, resulting in a reduced dielectric constant of the ferroelectric packing material. Thus, the electric field in the vicinity of the pellets is reduced resulting in reduced VOC removal efficiency.

3.2. Packing Material Size

The size of the packing materials plays an important role in determining the characteristics of discharges in the reactor and the efficiency of VOC decomposition. The diameter of the packing material is chosen to maximize the contact points between the packing materials of same size without completely filling the discharge gap. Ogata et al. [86,95] mentioned that BaTiO3 pellets of larger diameter (3 mm) are less efficient in decomposing benzene when compared to the smaller pellets. Similarly, Yamamoto et al. [38] mentioned that much higher toluene decomposition efficiency is obtained by using BaTiO3 pellets of smaller diameter. In [71], it is reported that BaTiO3 pellets of smaller diameter (1 mm) are more efficient in removing CCl4 when compared to pellets of diameter 3 mm. This is attributed to the enhanced electric field and more uniform plasma throughout the reactor volume when packed with smaller pellets. Contrarily, it has been reported in [96] that very fine titania pellets (~0.5 to 1 mm) reduced the removal efficiency of TCE. This is due to the insufficient generation of plasma in the reactive region as there is lack of void due to the compact packing of titania.

3.3. Packing Material Shape

Packing material made of the same material but with different shape can influence the capacitance of the reactor which changes the discharge characteristics [97] and thus VOC removal efficiency [79]. The waveforms of applied voltage and discharge current for various pellet shapes such as (a) sphere, (b) solid cylinder and (c) hollow cylinder are shown in Figure 6. The discharge current waveforms of the three different packing shapes show that the formation of microdischarges follows the following sequence: hollow cylinder > cylinder > sphere. Takaki et al. [90] showed that for the same applied voltage, the quantity of charge accumulated with the microdischarge current in a hollow cylindrical pellet is higher when compared to spherical and cylindrical pellet. The micro-discharges start appearing at lower voltage for the hollow cylinder. This is because the enhancement of electric field is higher for the hollow cylinder due to the presence of sharp edges [98]. The DBD reactor filled with hollow cylindrical pellet showed higher energy efficiency (g/kWh) in the abatement of C2F6 when compared with spherical pellets. In [96], Oda et al. showed that the removal efficiency showed by disc shaped randomly crushed TiO2 is higher when compared to spherical ones. This is due to the presence of sharp edges in the packing material which enhances the electric field.

3.4. Surface Properties

The surface properties of the packing material such as surface area, pore volume and pore size play an important role in complete oxidation of VOCs. These properties of the packing material affect the residence time of the VOCs in the discharge region of the reactor by adsorption and increases the collision probability between adsorbed pollutant molecules and the plasma generated active species. As the adsorbed molecules can only vibrate at their fixed position, they can be more easily attacked by the reactive species when compared to the randomly moving gas phase molecules. For example, adsorption packing materials such as Al2O3 and zeolites concentrate VOCs on their surface resulting in enhanced collisional probability and deep oxidation of VOCs. Adsorption lowers the bond energy of the VOC molecule [88] and weakens the chemical bond which enhances the dissociation when these adsorbed molecules encounters the active species in the plasma [99]. The most commonly used porous packing materials in DBD reactor for the abatement of VOCs are porous alumina, activated carbon and zeolites.
Lee et al. [100] reported that a reactor packed with porous γ-Al2O3 results in complete oxidation of benzene and enhanced CO2 selectivity due to an increase in residence time. Liang et al. [43] reported that Al2O3 increases the residence time of toluene in the reactor resulting in enhanced toluene removal rate. This is because porous alumina possess electron-accepting acidic center which acts as an adsorption material for compound with π-electrons [101]. He et al. [73] reported that the activated carbon exhibits higher removal efficiency of benzene due to its higher surface area (SBET = 609.3 m2/g) when compared to other packing materials such as alumina (SBET = 230 m2/g) and molecular sieve (SBET = 45 m2/g).
In literature, it has been reported that packing the reactor with the same material but different surface properties like surface area and porosity influences the oxidation of VOCs and formation of by-products. Gandhi et al. [102], reported that the ethylene decomposition efficiency is enhanced with porous alumina (SBET = 348.3 m2/g) when compared to non-porous alumina (SBET = 1.53 m2/g). Also the complete oxidation of ethylene to CO2 and CO is obtained with porous alumina, whereas non-porous alumina leads to the formation of by-products such as acetaldehyde, acetylene and methane in addition to CO and CO2. Ogata et al. [95] found that CO2 selectivity has been enhanced with a DBD reactor packed with porous Al2O3 when compared to non-porous Al2O3 due to complete oxidation of benzene in the adsorbed phase. Zheng et al. [88] reported that alumina with higher surface area (γ-Al2O3, SBET = 174 m2/g) shows better acetone removal due to adsorption when compared to the α-Al2O3 (SBET = 53 m2/g). In [42], it has been reported that the removal efficiency of formaldehyde (HCHO) with γ-Al2O3 (SBET ≈ 220 m2/g) is higher when compared to α-Al2O3 (SBET ≈ 10 m2/g) due to the difference in the specific surface area. Adsorption of HCHO reduces the dissociation threshold energy. Ogata et al. [44] reported that the energy efficiency of a Al2O3/BaTiO3 packed bed reactor is enhanced when compared to a BaTiO3 packed reactor. This finding shows that in addition to decomposition of benzene in the gas phase on the surface of BaTiO3, there is also the disappearance of benzene due to adsorption on porous Al2O3. In [96], it has been reported that titania pellets with higher surface area showed an enhanced removal efficiency of TCE. Pangilinan et al. [49] showed that silver loaded TiO2 with lower surface area (SBET = 49 m2/g) decomposes benzene up to 20%; whereas the Ag/TiO2 with higher surface area (SBET ≈ 63 m2/g) completely oxidizes benzene. The increase in surface area enhances the adsorption of benzene on a catalytic surface results in enhanced decomposition. In [18], Cu–Ce catalyst showed both a better removal efficiency of HCHO and a better CO2 selectivity when compared to CuO or CeO2. This is because the specific surface area and the pore volume of the Cu–Ce catalyst is larger when compared to the other catalysts which promotes the adsorption of VOCs and increases the retention time of the pollutant in the discharge region and enhances the deep oxidation [103].
Hu et al. [54] and Ogata et al. [104] reported that CO2 selectivity is greatly enhanced by zeolite hybrid reactors due to the porous structure of zeolites which increases the residence time of the adsorbed species resulting in an increased collisional probability between VOC and active species. The external surface area of zeolites is much smaller when compared to the total surface area (which includes micro and macro-pores). However the results suggest that benzene adsorbed on the external pores are readily oxidized when compared to the ones adsorbed on the internal pores [104]. The higher surface area and pore volume of HZSM-5 packing material enhance the adsorption of chlorobenzene and residence time and results in a higher removal efficiency and a good CO2 selectivity [105]. In [65], it has been reported that the presence of V2O5/TiO2 or Cu-ZSM-5 enhances the decomposition of TCE partly due to the adsorption of TCE on the packing material.
The porous packing material absorbs/adsorbs molecules of sizes smaller than its pore size [106]. When the pore size of a packing material is larger than the size of the VOC molecule, there is an increased probability of VOC adsorption in these micro-pores. In [41], benzene decomposition is studied using two types of zeolite, H-Y and ferrierrite, with pore diameter 7.4 Å and 4.3–5.3 Å respectively. It is shown that benzene (molecular size = 5.9 Å) can assimilate well in the micro-pores of H-Y when compared to ferrierite. This is because the pore size of H-Y is bigger than the molecular size of benzene.
Packing the DBD reactor with porous material (Al2O3 and SiO2) enhances the discharge power when compared to non-porous (ZrO2 and GW) packing material. This is due to the formation of micro-discharges in the micro-pores in addition to the micro-discharges in the gas phase [107]. In addition to the enhanced discharge power, the adsorption of ethylene in these porous material enhances the decomposition of ethylene, COx selectivity, carbon balance and reduces the formation of unwanted by-products such as N2O.
The effect of the presence of humidity in a gas stream on the removal of VOCs has to be considered for more realistic environmental and industrial applications. The presence of humidity in flue gasses can have both positive and negative effect on the decomposition of VOC and the product selectivity. The presence of humidity can modify the surface state of the packing material and quench the electrons which affects the discharge characteristics. The presence of water can greatly influence the chemical reactions by quenching electrons and other excited reactive species and by producing OH radicals.
Ogata et al. [46] reported that the presence of water (5%) reduces the decomposition efficiency of benzene, toluene and xylene in a BaTiO3 packed bed DBD reactor. Urashima et al. [92] showed that the removal efficiency of C2F6 is reduced in the presence of humidity (relative humidity = 70%) due to the extra electron energy loss for the dissociation of water molecules. In [47,108], it is reported that the presence of humidity changes the surface states of BaTiO3 in a PBDBD reactor, resulting in a reduced benzene removal efficiency. Wu et al. [68] reported that the presence of humidity (4.43 vol %) reduces the toluene removal efficiency due to quenching of high energy electrons by water and reducing the formation of active oxygen.
Zhu et al. [48] reported that the presence of humidity reduces the benzene removal efficiency due to the competitive adsorption of water on the surface of TiO2. Wu at al [68] reported that the presence of humidity reduces the toluene decomposition efficiency due to competitive adsorption of water on active sites of the catalyst. On the other hand, Zhao et al. [109] reported that the presence of humidity has no significant effect on the removal of formaldehyde using Ag/Cu-Hz as packing material due to the selective absorption. Thus, the humidity tolerance of the PBDBD reactor depends on the ability of packing to selectively adsorb the pollutant.
The presence of humidity suppresses the formation of toxic by-products such as CO and NOx and enhances CO2 selectivity. For the reactor packed with BaTiO3, the deactivation of lattice oxygen species in BaTiO3 by H2O reduces CO [46] and NOx formation [108].
The lifetime of the packing material is very important for the commercialization of the PBDBD reactors for the abatement of VOC. The micro-pores of the porous material can get clogged by the accumulation of solid deposits resulting in a decrease in the decomposition efficiency of VOCs. In [107], it is reported that the accumulation of solid deposit on the surface of porous packing material (α-Al2O3) blocks the pores leading to the deactivation of the packing material. The activity of the porous material can be regenerated by an oxygen plasma treatment of the packing material [107]. On the other hand, the non-porous packing material does not show any time dependent deterioration in its performance [107]. In [67], it has been reported that the presence of TiO2 enhances the benzene oxidation by decomposing ozone to form active oxygen. However, a poor carbon balance ( C B ( % ) = [ CO x ] 6 [ C 6 H 6 ] c o n v × 100 % ) is obtained due to deposition of intermediates on the surface of TiO2.

3.5. By-Product Formation

The production of toxic by-products during the abatement of VOC using NTP technology is one of the major concern for the commercialization of the discussed plasma technology for air purification. The most common by-products produced by an NTP reactor during the abatement of VOC in air are ozone, NOx, aerosols and other VOCs. It has been widely reported in literature that the presence of the packing material reduces the formation of toxic by-products such as ozone and NOx either by decomposition or by adsorption of the by-product and enhances CO2 selectivity by complete oxidation of organic compounds.
The fact that the production of ozone is reduced in the presence of a packing material is mainly due to the decomposition of ozone on the surface of the packing material to form active oxygen species. The activity of active oxygen is higher than ozone and this results in an enhanced VOC removal efficiency. Chae et al. [45] mentioned that the production of ozone in a PBDBD reactor is 60% lower than in an unpacked reactor. Zheng et al. [88] reported that the formation of ozone is highly reduced in the DBD with a packing material when compared to an empty reactor. In [110], ozone formation is suppressed by the presence of TiO2/Al2O3/Ni foam when compared to unpacked reactor. Whereas Zhu et al. [48] showed that the presence of TiO2 coated RR increases the ozone concentration enhancing benzene removal efficiency.
The CO2 selectivity during the abatement of VOCs is enhanced by the complete oxidation of the organic compounds to form CO2 and H2O. The formation of CO2 is enhanced by the presence of a porous packing material due to the adsorption of VOC in the pores which increases the VOC retention time in the discharge region resulting in deep oxidation of VOC. Ogata et al. [44,111] reported that the decomposition of benzene and CO2 selectivity are enhanced in an Al2O3/BaTiO3 packed reactor. Enhancement of CO2 selectivity has also been reported for reactors packed with photocatalysts such as TiO2/RR [48,53] and TiO2/glass beads [100]. Kim et al. [47] reported that a BaTiO3 packed bed reactor inhibits the formation of aerosol by enhancing the oxidation of benzene to CO and CO2.
Ogata et al. [44] found that the presence of porous Al2O3 in Al2O3/BaTiO3 in a hybrid PBDBD reactor suppresses the formation of N2O. Zheng et al. [88] reported that the formation of NO2 is highly reduced to 0.5 ppm in the DBD reactor packed with γ-Al2O3 when compared to an empty reactor. This is due to the adsorption of NO2 on the surface of the packing material and only 0.5 ppm of NO2 was produced.
The formation of yellow, brownish yellow or brownish solid deposit occurs on the wall of the DBD reactor which leads to a poor carbon balance during the abatement of VOC. The presence of packing material suppresses the formation of solid deposits by complete oxidation of the organic by-products. In [57], Gandhi et al. reported the formation of unwanted by-products such as acetylene and acetaldehyde from the decomposition of butane disappeared in the presence of glass wool loaded with TiO2 or GO as packing material. In [112], Trinh et al. reported that the presence of MnO2 coated monolith packing material inhibits the formation of polymer like deposit in the process of acetone abatement.
On the other hand, even in the presence of packing material, the formation of polymeric solid deposit is observed on the wall of the reactor and the surface of the packing material. During the decomposition of methane in the absence of O2, brown yellowish deposit was found in the inner electrode [113]. Decomposition of xylene in N2, BaTiO3 packed bed reactor results in poor carbon balance due to the formation of a carbon deposit, polymer, decomposed fragment, cluster, or particulate [46]. During the process of decomposition of benzene in the absence of O2, carbonaceous deposit was found on the surface of the catalyst [114,115]. Thus, it is evident that N2 plays an important role in the polymerization process and this problem can be rectified by using sufficient amount of O2 in the feed gas [57].

4. Catalyst Loading on Packing Materials

Catalysts in single stage reactors are in direct contact with plasma. The plasma assisted reactions on the surface of the catalyst play an important role in addition to the gas phase reaction in VOC removal [75]. The pathway of destruction of one of the model VOC (i.e., Toluene) in FeOx/SBA-15 packed bed DBD reactor is shown in Figure 7. As shown in the mechanism pathway, the destruction of toluene takes place: (i) in gas phase (by direct collision of toluene and other intermediate by-products with electron and other reactive species such as O•, OH•, N2•, NO•, N2O•) and (ii) on the surface of the catalyst (by the reaction between absorbed toluene and other intermediate by-products with active species such as O• and OH•) [116]. The combined effect of the excellent toluene absorption capability of SBA-15 and O2 absorption capacity of Fe2+ greatly increased the toluene removal efficiency, COx selectivity and remarkably reduced the formation organic byproducts. Thus, the key points for screening the catalyst for the abatement of VOCs are removal efficiency and carbon balance. In literature, it is shown that CO2 selectivity is independent of SIE whereas the removal efficiency increases with increase in SIE. The initial decomposition of VOC is determined by the SIE applied, whereas the catalyst is crucial in the decomposition of the intermediates formed during the decomposition of VOC resulting in a better carbon balance. Thus, the catalyst loading on the packing materials improves the VOC removal efficiency [73], suppresses the formation of unwanted by-products like CO, NOx [73], ozone [117] by deep oxidation of the intermediate by-products. The most commonly used catalyst in the PBDBD reactors for the abatement of VOCs are metal oxides such as CuOx [18,68,73,117,118], MnOx [67,68,73,81,105,117,119,120,121], CeOx [18,68,74,105,117,122], CoOx [117], NiOx [68,117], FeOx [116] and noble metal coated or impregnated porous or metal oxide catalysts.

4.1. Removal Efficiency

The transition metal oxides have lattice oxygen, surface oxygen and adsorbed oxygen [123,124] which play an important role in VOC oxidation in addition to the gas phase oxygen present in the flue gas [125,126]. Zhu et al. [118] reported that a better acetone removal efficiency was obtained for 5 wt % CuO/γ-Al2O3 due to its higher reducibility and abundance of highly mobile oxygen species on the surface of the catalyst. In [18], it has been reported that the combination of plasma and Cu–Ce binary metal catalyst enhanced the removal efficiency of formaldehyde. This is mainly attributed to the enhanced oxygen vacancy and adsorbed oxygen on the pores of the surface of the catalyst which facilitate the redox reaction between Ce and Cu species in the catalyst. Zhu et al. [53] reported that the decomposition efficiency of benzene is improved by the presence of TiO2 catalyst in a RR packed bed DBD reactor. In [121], it has been reported that WO3 coated on TiO2 has better removal efficiency of TCE when compared to V2O5 coated TiO2; however the reason for this behavior is not explained in the article. In [96], it has been reported that the removal efficiency of TCE is enhanced by coating TiO2 with V2O5.
Uniform dispersion of a metal catalyst on the packing material is very important for enhanced VOC removal efficiency due to the increase in the active metal surface [49]. Lu et al. [116] reported that loading SBA-15 with FeOx enhances the toluene removal efficiency and CO2 selectivity by reducing the unwanted by-products like benzaldehyde, benzyl alcohol, formic acid, acetic acid, benzenemethanol, tetradecane, hexadecane and heptadecane. The excellent toluene absorbing property of SBA-15 and oxygen absorbing property of Fe2+ ion generated by FeOx are decisive in the abatement of toluene. 3% Fe loading is found to be optimal due to its high dispersion nature on the support material. In [68], it is shown that NiO loaded γ-Al2O3 enhanced the removal of toluene as NiO decomposes ozone to atomic oxygen much more efficiently when compared to other catalysts. The metal loading up to 5 wt % increases the removal efficiency whereas further increase in Ni loading decreases the decomposition efficiency. This is probably due to the formation of bulk NiO which starts screening the active sites as 5 wt % of Ni loading is closer to the threshold of monolayer dispersion of NiO on Al2O3. Kang et al. [127] reported that TiO2 supported by Al2O3 shows better decomposition efficiency of toluene when compared to glass beads. This is because more TiO2 is attached to Al2O3 than to glass beads and Al2O3 acts as an electron acceptor.
On the other hand, loading the porous support material with catalyst can affect the surface properties of the packing material by reducing the overall pore volume and surface area. Zhu et al. [117] reported that the removal efficiency of acetone is enhanced with MOx/γ-Al2O3 when compared to pure γ-Al2O3 even though the metal oxide supported alumina has smaller surface area (SBET = 148 ± 2 m2/g) when compared to pure alumina (SBET = 187 m2/g). This emphasizes the role of catalytic activity of the metal oxides in the decomposition of VOCs. The catalytic activity of metal oxides is determined by their reducibility and the presence of oxygen on the surface of the catalyst. The acetone removal efficiency of CuOx/γ-Al2O3 is higher when compared to other metal oxides due to its higher reducibility and a carbon balance ( C B ( % ) = 3 [ C 3 H 6 O ] o u t + [ CO ] + [ HCOOH ] + [ CO ] 2 + [ HCHO ] 3 [ C 3 H 6 O ] i n × 100 % ) higher than 95% is obtained due to the presence of surface oxygen species.
The presence of an oxidation catalyst as packing material plays an important role in the formation of active oxygen on the surface of the catalyst which enhances VOC removal efficiency. The active oxygen is formed from ozone by two different pathways as follows: (i) direct collision between active species and air molecules (Equations (4) and (5)) [105] and (ii) decomposition of ozone on the catalyst surface (Equation (6)). The amount of active atomic oxygen produced by the second pathway is dependent on the nature of the catalyst. In [119], it has been reported that the presence of MnO2 enhances the decomposition of TCE and reduced the formation of by-products such as dichloro-acetylchloride (DCAC), phosgene and trichloro-acetaldehyde (TCAA). In [67], it has been reported that the presence of MnO2 packing material enhances the decomposition of benzene. In [128], it is shown that the MnO2 loaded γ-Al2O3 enhances the removal of toluene due to an enhancement of the redox reaction by MnO2. In [105], it has been shown that CuO supported on MnO2 enhances the removal efficiency of chlorobenzene. Thus, MnO2 accelerates the decomposition of ozone to form active O species resulting in an enhanced and complete oxidation of VOCs.
e + O 2 2 O + e
e + O 3 O + O 2 + e
O 3 + *   O 2 + O *
The removal efficiency of chlorobenzene and CO2 selectivity are enhanced by a CeO2/HZSM-5 packed reactor when compared with an unpacked reactor due to the good catalytic activity, the presence of oxygen vacancy and the oxygen activity of CeO2 [105] and due to the high surface area and Brønsted acid sites of HZSM-5 [122]. The level of ozone in the exhaust is also reduced, probably due to the decomposition of ozone on the surface of the catalyst [129] and utilized for chlorobenzene conversion. In [65], it has been reported that the presence of V2O5/TiO2 or Cu-ZSM-5 enhances the decomposition of TCE partly due to the adsorption of TCE on the packing material.
Kim et al. [41] studied the effect of Ag loading on TiO2 on the decomposition efficiency of benzene and the carbon balance. The results showed that higher the silver content the better the carbon balance ( C B ( % ) = [ CO ] + [ CO ] 2 + [ HCOOH ] n [ VOC ] c o n v × 100 % ) is as silver aids the decomposition of intermediates formed on the surface of TiO2. However the decomposition efficiency starts to decline when the Ag loading is more than 4 wt %, as the loaded silver decreases the surface area of TiO2 which is responsible for the initial decomposition of benzene. Whereas in [120], it is mentioned that the TiO2 is more effective in benzene decomposition when compared to 0.5 wt % Ag impregnated TiO2. In the above mentioned works, although the Ag loading on TiO2 is done by the impregnation method, the difference in the results can be explained by the difference in calcination (thermal treatment of catalyst in oxidation environment) temperature and calcination time of the catalyst. Kim et al. [41] studied the loading of different metals such as Ag and Ni on the TiO2 packing material for the removal efficiency of benzene and carbon balance. Although Ni loaded TiO2 showed higher removal efficiencies when compared to Ag, there was a problem of a poor carbon balance. This shows that Ni has lower catalytic reactivity towards the intermediates produced during the process and these intermediates are deposited on the surface of Ni/TiO2.
Li et al. [50] studied the effect of Pt loading on γ-Al2O3 on the decomposition efficiency of benzene, CO2 selectivity and formation of by-products. Pt/γ-Al2O3 showed higher decomposition efficiency when compared to unloaded γ-Al2O3. This is because the presence of Pt increases the number of active sites and reduces the activation energy of the decomposition reaction. Pt/γ-Al2O3 also suppresses harmful NOx formation by (i) reducing the energy of electrons; thus reducing the number of excited N2, (ii) accelerate oxygen decomposition and oxidation. Also, the formation of CO is reduced in the presence of Pt/γ-Al2O3 as Pt can promote O3 decomposition resulting in enhanced CO oxidation. An et al. [62] studied the effect of metal (Cu) loading on molecular sieve and Cu, Co, Ag or Au loading on Al2O3 on the removal efficiency of toluene and formation of by-products and the results show that the metal loading enhances the removal efficiency by decomposition of ozone on the catalyst surface.

4.2. By-Product Formation

The formation of unwanted by-products such as CO, ozone, NOx and other organic intermediates is one of the main disadvantages of NTP application for the abatement of VOCs. In the literature, it has been widely discussed that the metal catalysts loaded packing material plays an important role in the deep oxidation of VOCs resulting in increased CO2 selectivity due to deep oxidation of the intermediate organic by-products.
The formation of CO is reduced by using Ni loaded BaTiO3 as packing material when compared to a BaTiO3 pellet packed bed reactor [71]. Apart from the enhanced removal efficiency, Pt-Pd loaded Al2O3 drive the oxidation of styrene towards CO2 rather than the formation of CO [75]. In [51], it is shown that loading HZSM-5 with Ag results in complete oxidation of benzene to CO2 and only a very negligible amount of CO is produced. This is due to strong π-complexation of Ag with benzene [101]. Kim et al. [55] reported that the catalyst loading (Ag/Pt) on TiO2 enhances CO2 selectivity during benzene decomposition. However Pt loaded TiO2 has a negative effect on the decomposition efficiency when compared to TiO2. In [57], it is reported that glass wool loaded with TiO2 or graphene oxide produced less CO and other organic by-products when compared to glass wool loaded with ZrO2 or alumina. This is probably due to the photocatalytic properties of graphene oxide and TiO2 which enhanced the formation of CO2.
The amount of catalyst loading on the packing material plays an important role in CO2 selectivity. Kim et al. [52] reported that CO2 selectivity is independent of SIE when Ag loading is more than 0.5 wt %; whereas for Ag loading less than 0.3 wt %, CO2 selectivity decreases with increase in SIE due to the deposit of intermediates on the packing material.
Ogata et al. [95] mentioned that the formation of N2O during decomposition of benzene is greatly reduced by packing the DBD reactor with metal loaded porous Al2O3/BaTiO3. Kim et al. [55] reported that the catalyst (Ag/Pt) loaded TiO2 suppresses the formation of undesired by-product such as NOx. An et al. [62] reported that NOx produced by the plasma discharge during toluene decomposition is reduced by the presence of 1 wt % Au/Al2O3 as packing material. Pangilinan et al. [49] reported that the formation of ozone is reduced with Ag/TiO2 and decreases with the increase in metal loading on the catalyst.

4.3. Catalyst Poisoning

The activity of the catalyst used as packing material in PBDBD reactor shows time dependent deterioration due to the accumulation of organic intermediates on the surface of the catalyst or absorption of certain organic compounds.
During the process of decomposition of benzene, toluene and xylene, the activity of MnOx/Al2O3 shows a decreased conversion efficiency after 150 h of usage [130]. This is probably due to the formation of the organic intermediates which is mainly composed of saturated aldehydes and/or ketones, carboxylic acids, phenols, primary alcohols, acid amides and nitro-aromatic compounds on the surface of the catalysts [130]. In [122], Jiang et al. reported the deactivation of catalyst in CeO2/HZSM-5 packing material which is used for decomposition of chlorobenzene after 75 h of usage. The deactivation is mainly due to the absorption and deposition of methyl and organic halide on the surface of the catalyst which covers the active sites of the catalyst. In [131], Ran et al. studied the catalytic activity of Ru/Al2O3 packing material used for the abatement of dichloromethane in PBDBD reactor and observed the deactivation of catalysts which is mainly due to the chlorine deposition. Zhu et al. reported slight deactivation of Cu1Ce1 catalyst after 5 h of usage for the abatement of formaldehyde [18]. This is due to the formation of coke on the surface of the catalyst which blocks the micropores [18].
As prolonging the service life of the catalyst is very important for practical application, a suitable method for the regeneration of deactivated catalyst is very important. In most of the cases, the activity of the catalyst is gained by calcination of the catalyst in the presence of O2 or air flow. In [130], Fan et al. gained the activity of MnOx/Al2O3 by calcination of catalyst in air stream at 773 K for 4 h and decomposition efficiency obtained by using the calcined catalyst showed that the catalyst was fully regenerated.

5. Conclusions

In this paper, an overview of the literature on packed bed dielectric barrier discharge (PBDBD) reactors for the abatement of volatile organic compounds is presented. The configuration of the packed bed DBD reactors such as the powered electrode, ground electrode and the power supply play an important role in determining the energy efficiency of the reactor. Using bolt as high voltage electrode enhances the electric field due to the presence of sharp edges and increases the number of microdischarges. Silver paint as ground electrode is advantageous as it avoids the formation of corona discharge in the voids outside the reactor. Thus, bolt as high voltage electrode and silver paint as ground electrode enhances the energy efficiency of the PBDBD reactors.
From this review, it is evident that the VOC removal efficiency of packed bed reactors is higher than of unpacked reactors. The presence of packing material in the reactor enhances the electric field in the area near the contact points between the pellets and homogenously distributes the plasma in the reactor. The effect of various properties of packing materials such as dielectric constant, size, shape, surface properties (surface area, pore volume and pore size) on the discharge characteristics and abatement of different VOCs has been reviewed. The dielectric constant of the packing material is important in determining the discharge characteristics and decomposition efficiency of the PBDBD. The packing material with higher dielectric constant increases the decomposition efficiency of the reactor and the increase in decomposition efficiency saturates at particular dielectric constant of the packing material. The size of the packing material is chosen to maximize the contact points between the packing material without completely filling the reactor. Maximizing the contact points enhances the electric field and produces uniform plasma and the presence of gaps in the packing facilitates the generation of plasma. The hollow cylindrical shape of the packing material enhances the electric field due to the presence of sharp edges and thus increases the decomposition efficiency of the PBDBD reactor. The presence of a porous packing material increases the residence time of VOC in the discharge zone by adsorption and enhances the collisional probability between a VOC molecule and plasma generated active species. Porous material with higher surface area enhances the decomposition efficiency, increases CO2 selectivity and results in complete oxidation of VOC. The pore size of the packing material is crucial for the adsorption of VOC as the porous material is efficient in adsorbing molecules smaller than its pore size. The porous packing material also enhances the discharge power due to the formation of additional microdischarges inside the pores. On the other hand, porous packing material shows deactivation with continuous usage due to the formation of coke deposit on the surface which blocks the pores. The activity of the packing material can be regenerated by oxygen plasma treatment.
Apart from enhancing the energy efficiency and decomposition efficiency, the presence of packing material significantly reduces the formation of unwanted by-products and enhances CO2 selectivity. The presence of packing material reduces the formation of ozone due to decomposition of ozone on the surface of packing material. Packing the DBD reactor with porous packing material increases CO2 selectivity and suppresses NOx formation. On the other hand, the problem of formation of polymeric deposit due to continuous usage is still prevalent in the packed bed DBD reactor which results in poor carbon balances. The formation of polymeric deposit can be significantly reduced by presence of at least 1% O2 in the feed gas and the solid carbon deposit can be removed using oxygen plasma.
The presence of a catalytic packing material produces active oxygen on the surface of the catalyst by the decomposition of ozone and this enhances the deep oxidation of VOCs and reduces the ozone concentration in the exhaust. The amount of metal loading on the packing material plays an important role in the formation of CO2 by deep oxidation of organic intermediate products. The presence of metal catalysts in the packing material suppresses the formation of oxides of nitrogen which is one of the important toxic by-products produced in the presence of air in the feed gas.
Thus, the presence of packing material in the discharge region of the dielectric barrier discharge reactor enhances the energy efficiency, increases VOC removal efficiency, suppresses the formation of unwanted toxic by-products such as NOx, ozone, other VOCs and significantly improves CO2 selectivity when compared to the unpacked DBD reactor. On the other hand, for the commercialization of this technology, the lifetime of the porous packing material, deactivation of catalyst and the reaction kinetics for the different types of pollutants and catalysts in PBDBD reactors are important for this subject of research. An extensive investigation on the modification of the packing material and catalyst used in packed bed DBD reactors by the non-thermal plasma will be an important subject for this field of research.

Acknowledgments

This research has received funding from the Interreg 2014-2020 France-Wallonie-Vlaanderen program (EFRO) through the Depollutair project.

Author Contributions

Savita K. P. Veerapandian wrote the first draft of the article which was then refined by the comments and suggestions from Christophe Leys, Nathalie De Geyter and Rino Morent.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nikolajsen, K.; Kiwi-Minsker, L.; Renken, A. Structured fixed-bed adsorber based on zeolite/sintered metal fibre for low concentration VOC removal. Chem. Eng. Res. Des. 2006, 84, 562–568. [Google Scholar] [CrossRef]
  2. Ozturk, B.; Yilmaz, D. Absorptive removal of volatile organic compounds from flue gas streams. Process Saf. Environ. Prot. 2006, 84, 391–398. [Google Scholar] [CrossRef]
  3. Poddar, T.K.; Majumdar, S.; Sirkar, K.K. Removal of VOCs from air by membrane-based absorption and stripping. J. Memb. Sci. 1996, 120, 221–237. [Google Scholar] [CrossRef]
  4. Dwivedi, P.; Gaur, V.; Sharma, A.; Verma, N. Comparative study of removal of volatile organic compounds by cryogenic condensation and adsorption by activated carbon fiber. Sep. Purif. Technol. 2004, 39, 23–37. [Google Scholar] [CrossRef]
  5. Gupta, V.K.; Verma, N. Removal of volatile organic compounds by cryogenic condensation followed by adsorption. Chem. Eng. Sci. 2002, 57, 2679–2696. [Google Scholar] [CrossRef]
  6. Li, W.; Gong, H. Recent progress in the removal of volatile organic compounds by catalytic combustion. Acta Phys. Chim. Sin. 2010, 26, 885–894. [Google Scholar]
  7. Qu, Z.; Bu, Y.; Qin, Y.; Wang, Y.; Fu, Q. The effects of alkali metal on structure of manganese oxide supported on SBA-15 for application in the toluene catalytic oxidation. Chem. Eng. J. 2012, 209, 163–169. [Google Scholar] [CrossRef]
  8. Demeestere, K.; Dewulf, J.; Van Langenhove, H. Heterogeneous photocatalysis as an advanced oxidation process for the abatement of chlorinated, monocyclic aromatic and sulfurous volatile organic compounds in air: State of the art. Crit. Rev. Environ. Sci. Technol. 2007, 37, 489–538. [Google Scholar] [CrossRef]
  9. Einaga, H.; Ibusuki, T.; Futamura, S. Photocatalytic oxidation of benzene in air. J. Sol. Energy Eng. 2004, 126, 789–793. [Google Scholar] [CrossRef]
  10. Mo, J.; Zhang, Y.; Xu, Q.; Lamson, J.J.; Zhao, R. Photocatalytic purification of volatile organic compounds in indoor air: A literature review. Atmos. Environ. 2009, 43, 2229–2246. [Google Scholar] [CrossRef]
  11. Kumar, A.; Dewulf, J.; Van Langenhove, H. Membrane-based biological waste gas treatment. Chem. Eng. J. 2008, 136, 82–91. [Google Scholar] [CrossRef]
  12. Mudliar, S.; Giri, B.; Padoley, K.; Satpute, D.; Dixit, R.; Bhatt, P.; Pandey, R.; Juwarkar, A.; Vaidya, A. Bioreactors for treatment of VOCs and odours—A review. J. Environ. Manag. 2010, 91, 1039–1054. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, X.; Jia, X.; Wen, J. Transient modeling of toluene waste gas biotreatment in a gas-liquid airlift loop reactor. Chem. Eng. J. 2010, 159, 1–10. [Google Scholar] [CrossRef]
  14. Khan, F.I.; Kr. Ghoshal, A. Removal of volatile organic compounds from polluted air. J. Loss Prev. Process Ind. 2000, 13, 527–545. [Google Scholar] [CrossRef]
  15. Rani Parmar, G.; Rao, N.N. Emerging control technologies for volatile organic compounds. Crit. Rev. Environ. Sci. Technol. 2009, 39, 41–78. [Google Scholar] [CrossRef]
  16. Luengas, A.; Barona, A.; Hort, C.; Gallastegui, G.; Platel, V.; Elias, A. A review of indoor air treatment technologies. Rev. Environ. Sci. Biotechnol. 2015, 14, 499–522. [Google Scholar] [CrossRef]
  17. Hermia, J.; Vigneron, S. Catalytic incineration for odour abatement and VOC destruction. Catal. Today 1993, 17, 349–358. [Google Scholar] [CrossRef]
  18. Zhu, X.; Gao, X.; Qin, R.; Zeng, Y.; Qu, R.; Zheng, C.; Tu, X. Plasma-catalytic removal of formaldehyde over Cu–Ce catalysts in a dielectric barrier discharge reactor. Appl. Catal. B 2015, 170–171, 293–300. [Google Scholar] [CrossRef]
  19. Sivachandiran, L.; Karuppiah, J.; Subrahmanyam, C. DBD plasma reactor for oxidative decomposition of chlorobenzene. Int. J. Chem. React. Eng. 2012. [Google Scholar] [CrossRef]
  20. Chen, H.L.; Lee, H.M.; Chen, S.H.; Chang, M.B.; Yu, S.J.; Li, S.N. Removal of volatile organic compounds by single-stage and two-stage plasma catalysis systems: A review of the performance enhancement mechanisms, current status, and suitable applications. Environ. Sci. Technol. 2009, 43, 2216–2227. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, H.-H. Nonthermal plasma processing for air-pollution control: A historical review, current issues, and future prospects. Plasma Process. Polym. 2004, 1, 91–110. [Google Scholar] [CrossRef]
  22. Vandenbroucke, A.M.; Morent, R.; De Geyter, N.; Leys, C. Non-thermal plasmas for non-catalytic and catalytic VOC abatement. J. Hazard. Mater. 2011, 195, 30–54. [Google Scholar] [CrossRef] [PubMed]
  23. Oda, T. Non-thermal plasma processing for environmental proection: Decomposition of dilute VOCs in air. J. Electrostat. 2003, 57, 293–311. [Google Scholar] [CrossRef]
  24. Van Durme, J.; Dewulf, J.; Leys, C.; Van Langenhove, H. Combining non-thermal plasma with heterogeneous catalysis in waste gas treatment: A review. Appl. Catal. B 2008, 78, 324–333. [Google Scholar] [CrossRef]
  25. Trinh, Q.H.; Mok, Y.S. Effect of the adsorbent/catalyst preparation method and plasma reactor configuration on the removal of dilute ethylene from air stream. Catal. Today 2014, 256, 170–177. [Google Scholar] [CrossRef]
  26. Li, S.; Tang, Z.C.; Gu, F. Experimental study on temperature characteristics and energy conversion in packed bed reactor with dielectric barrier discharge. Heat Mass Transf. 2010, 46, 851–857. [Google Scholar] [CrossRef]
  27. Huang, L.; Nakajo, K.; Ozawa, S.; Matsuda, H. Decomposition of dichloromethane in a wire-in-tube pulsed corona reactor. Environ. Sci. Technol. 2001, 35, 1276–1281. [Google Scholar] [CrossRef] [PubMed]
  28. Penetrante, B.M.; Hsiao, M.C.; Bardsley, J.N.; Merritt, B.T.; Vogtlin, G.E.; Kuthi, A.; Burkhart, C.P.; Bayless, J.R. Decomposition of methylene chloride by electron beam and pulsed corona processing. Phys. Lett. A 1997, 235, 76–82. [Google Scholar] [CrossRef]
  29. Hsiao, M.C.; Merritt, B.T.; Penetrante, B.M.; Vogtlin, G.E.; Wallman, P.H. Plasma-assisted decomposition of methanol and trichloroethylene in atmospheric pressure air streams by electrical discharge processing. J. Appl. Phys. 1995, 78, 3451–3456. [Google Scholar] [CrossRef]
  30. Penetrante, B.M.; Hsiao, M.C.; Bardsley, J.N.; Merritt, B.T.; Vogtlin, G.E.; Wallman, P.H.; Kuthi, A.; Burkhart, C.P.; Bayless, J.R. Electron beam and pulsed corona processing of carbon tetrachloride in atmospheric pressure gas streams. Phys. Lett. A 1995, 209, 69–77. [Google Scholar] [CrossRef]
  31. Norberg, A. Modeling current pulse shape and energy in surface discharges. IEEE Trans. Ind. Appl. 1992, 28, 498–503. [Google Scholar] [CrossRef]
  32. Rousseau, A.; Dantier, A.; Gatilova, L.; Ionikh, Y.; Röpcke, J.; Tolmachev, Y. On NOx production and volatile organic compound removal in a pulsed microwave discharge in air. Plasma Sources Sci. Technol. 2005, 14, 70–75. [Google Scholar] [CrossRef]
  33. Kogelschatz, U. Dielectric-barrier discharges: Their history, discharge physics, and industrial applications. Plasma Chem. Plasma Process. 2003, 23, 1–46. [Google Scholar] [CrossRef]
  34. Evans, D.; Rosocha, L.A.; Anderson, G.K.; Coogan, J.J.; Kushner, M.J. Plasma remediation of trichloroethylene in silent discharge plasmas. J. Appl. Phys. 1993, 74, 5378–5386. [Google Scholar] [CrossRef]
  35. Falkenstein, Z. Effects of the O2 concentration on the removal efficiency of volatile organic compounds with dielectric barrier discharges in Ar and N2. J. Appl. Phys. 1999, 85, 525–529. [Google Scholar] [CrossRef]
  36. Snyder, H.R.; Anderson, G.K. Effect of air and oxygen content on the dielectric barrier discharge decomposition of chlorobenzene. IEEE Trans. Plasma Sci. 1998, 26, 1695–1699. [Google Scholar] [CrossRef]
  37. Oda, T.; Takahashi, T.; Tada, K. Decomposition of dilute trichloroethylene by nonthermal plasma. IEEE Trans. Ind. Appl. 1999, 35, 373–379. [Google Scholar] [CrossRef]
  38. Yamamoto, T.; Ramanathan, K.; Lawless, P.A.; Ensor, D.S.; Newsome, J.R.; Plaks, N.; Ramsey, G.H. Control of volatile organic compounds by an ac energized ferroelectric pellet reactor and a pulsed corona reactor. IEEE Trans. Ind. Appl. 1992, 28, 528–534. [Google Scholar] [CrossRef]
  39. Tonkyn, R.G.; Barlow, S.E.; Orlando, T.M. Destruction of carbon tetrachloride in a dielectric barrier/packed-bed corona reactor. J. Appl. Phys. 1996, 80, 4877–4886. [Google Scholar] [CrossRef]
  40. Yamamoto, T. Optimization of nonthermal plasma for the treatment of gas streams. J. Hazard. Mater. 1999, 67, 165–181. [Google Scholar] [CrossRef]
  41. Kim, H.H.; Ogata, A.; Futamura, S. Effect of different catalysts on the decomposition of VOCs using flow-type plasma-driven catalysis. IEEE Trans. Plasma Sci. 2006, 34, 984–995. [Google Scholar] [CrossRef]
  42. Ding, H.-X.; Zhu, A.-M.; Yang, X.-F.; Li, C.-H.; Xu, Y. Removal of formaldehyde from gas streams via packed-bed dielectric barrier discharge plasmas. J. Phys. D 2005, 38, 4160–4167. [Google Scholar] [CrossRef]
  43. Liang, W.J.; Wang, A.; Ma, L.; Li, J. Combination of spontaneous polarization plasma and photocatalyst for toluene oxidation. J. Electrostat. 2015, 75, 27–34. [Google Scholar] [CrossRef]
  44. Ogata, A.; Yamanouchi, K.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Oxidation of dilute benzene in an alumina hybrid plasma reactor at atmospheric pressure. Plasma Chem. Plasma Process. 1999, 19, 383–394. [Google Scholar] [CrossRef]
  45. Chae, J.; Moon, S.; Sun, H.; Kim, K.; Vassiliev, V.A.; Mikholap, E.M. A study of volatile organic compounds decomposition with the use of non-thermal plasma. KSME Int. J. 1999, 13, 647–655. [Google Scholar] [CrossRef]
  46. Ogata, A.; Ito, D.; Mizuno, K.; Kushiyama, S.; Gal, A.; Yamamoto, T. Effect of coexisting components on aromatic decomposition in a packed-bed plasma reactor. Appl. Catal. A 2002, 236, 9–15. [Google Scholar] [CrossRef]
  47. Kim, H.H.; Kobara, H.; Ogata, A.; Futamura, S. Comparative assessment of different nonthermal plasma reactors on energy efficiency and aerosol formation from the decomposition of gas-phase benzene. IEEE Trans. Ind. Appl. 2005, 41, 206–214. [Google Scholar] [CrossRef]
  48. Zhu, T.; Li, J.; Jin, Y.; Liang, Y.; Ma, G. Decomposition of benzene by non-thermal plasma processing: Photocatalyst and ozone effect. Int. J. Environ. Sci. Technol. 2008, 5, 375–384. [Google Scholar] [CrossRef]
  49. Pangilinan, C.D.C.; Kurniawan, W.; Salim, C.; Hinode, H. Effect of Ag/TiO2 catalyst preparation on gas-phase benzene decomposition using non-thermal plasma driven catalysis under oxygen plasma. React. Kinet. Mech. Catal. 2016, 117, 103–118. [Google Scholar] [CrossRef]
  50. Li, J.; Han, S.; Bai, S.; Han, S.; Song, H.; Pu, Y.; Zhu, X.; Chen, W. Effect of Pt/gamma-Al2O3 catalyst on nonthermal plasma decomposition of benzene and byproducts. Environ. Eng. Sci. 2011, 28, 395–403. [Google Scholar] [CrossRef]
  51. Fan, H.-Y.; Shi, C.; Li, X.-S.; Zhao, D.-Z.; Xu, Y.; Zhu, A.-M. High-efficiency plasma catalytic removal of dilute benzene from air. J. Phys. D 2009, 42, 225105. [Google Scholar] [CrossRef]
  52. Kim, H.H.; Oh, S.M.; Ogata, A.; Futamura, S. Decomposition of benzene using Ag/TiO2 packed plasma-driven catalyst reactor: Influence of electrode configuration and Ag-loading amount. Catal. Lett. 2004, 96, 189–194. [Google Scholar] [CrossRef]
  53. Zhu, T.; Li, J.; Jin, Y.Q.; Liang, Y.H.; Ma, G.D. Gaseous phase benzene decomposition by non-thermal plasma coupled with nano titania catalyst. Int. J. Environ. Sci. Technol. 2009, 6, 141–148. [Google Scholar] [CrossRef]
  54. Hu, J.; Jiang, N.; Li, J.; Shang, K.; Lu, N.; Wu, Y. Degradation of benzene by bipolar pulsed series surface/packed-bed discharge reactor over MnO2-TiO2/zeolite catalyst. Chem. Eng. J. 2016, 293, 216–224. [Google Scholar] [CrossRef]
  55. Kim, H.H.; Lee, Y.H.; Ogata, A.; Futamura, S. Plasma-driven catalyst processing packed with photocatalyst for gas-phase benzene decomposition. Catal. Commun. 2003, 4, 347–351. [Google Scholar] [CrossRef]
  56. Jiang, N.; Lu, N.; Li, J.; Wu, Y. Degradation of benzene by using a silent-packed bed hybrid discharge plasma reactor. Plasma Sci. Technol. 2012, 14, 140–146. [Google Scholar] [CrossRef]
  57. Gandhi, M.S.; Mok, Y.S.; Lee, S.B.; Park, H. Effect of various parameters for butane decomposition under ambient temperature in a dielectric barrier discharge non-thermal plasma reactor. J. Taiwan Inst. Chem. Eng. 2013, 44, 786–794. [Google Scholar] [CrossRef]
  58. Delagrange, S.; Pinard, L.; Tatibout, J.M. Combination of a non-thermal plasma and a catalyst for toluene removal from air: Manganese based oxide catalysts. Appl. Catal. B 2006, 68, 92–98. [Google Scholar] [CrossRef]
  59. Mok, Y.S.; Nam, C.M.; Cho, M.H.; Nam, I.S. Decomposition of volatile organic compounds and nitric oxide by nonthermal plasma discharge processes. IEEE Trans. Plasma Sci. 2002, 30, 408–416. [Google Scholar]
  60. Lee, H.M.; Chang, M.B. Gas-phase removal of acetaldehyde via packed-bed dielectric barrier discharge reactor. Plasma Chem. Plasma Process. 2001, 21, 329–343. [Google Scholar] [CrossRef]
  61. Liang, W.; Li, J.; Li, J.; Jin, Y. Abatement of toluene from gas streams via ferro-electric packed bed dielectric barrier discharge plasma. J. Hazard. Mater. 2009, 170, 633–638. [Google Scholar] [CrossRef] [PubMed]
  62. Quoc An, H.T.; Pham Huu, T.; Le Van, T.; Cormier, J.M.; Khacef, A. Application of atmospheric non thermal plasma-catalysis hybrid system for air pollution control: Toluene removal. Catal. Today 2011, 176, 474–477. [Google Scholar] [CrossRef]
  63. Dou, B.; Bin, F.; Wang, C.; Jia, Q.; Li, J. Discharge characteristics and abatement of volatile organic compounds using plasma reactor packed with ceramic Raschig rings. J. Electrostat. 2013, 71, 939–944. [Google Scholar] [CrossRef]
  64. Takaki, K.; Hatanaka, Y.; Arima, K.; Mukaigawa, S.; Fujiwara, T. Influence of electrode configuration on ozone synthesis and microdischarge property in dielectric barrier discharge reactor. Vacuum 2008, 83, 128–132. [Google Scholar] [CrossRef]
  65. Oda, T.; Yamaji, K.; Takahashi, T. Decomposition of dilute trichloroethyle by non-thermal plasma processing-frequency and catalyst effects. IEEE Trans. Ind. Appl. 2001, 37, 965–970. [Google Scholar] [CrossRef]
  66. Ogata, A.; Miyamae, K.; Mizuno, K.; Kushiyama, S.; Tezuka, M. Decomposition of benzene in air in a plasma reactor: Effect of reactor type and operating conditions. Plasma Chem. Plasma Process. 2002, 22, 537–552. [Google Scholar] [CrossRef]
  67. Futamura, S.; Einaga, H.; Kabashima, H.; Hwan, L.Y. Synergistic effect of silent discharge plasma and catalysts on benzene decomposition. Catal. Today 2004, 89, 89–95. [Google Scholar] [CrossRef]
  68. Wu, J.; Huang, Y.; Xia, Q.; Li, Z. Decomposition of toluene in a plasma catalysis system with NiO, MnO2, CeO2, Fe2O3, and CuO catalysts. Plasma Chem. Plasma Process. 2013, 33, 1073–1082. [Google Scholar] [CrossRef]
  69. Kim, H.H.; Oh, S.M.; Ogata, A.; Futamura, S. Decomposition of gas-phase benzene using plasma-driven catalyst (PDC) reactor packed with Ag/TiO2 catalyst. Appl. Catal. B 2005, 56, 213–220. [Google Scholar] [CrossRef]
  70. Futamura, S.; Yamamoto, T.; Lawless, P.A. Towards understanding of VOC decomposition mechanisms using nonthermal plasmas. In Proceedings of the 1995 Thirtieth IAS Annual Meeting IEEE Conference Record of Industry Applications Conference, Orlando, FL, USA, 8–12 October 1995; Volume 2, pp. 1453–1458. [Google Scholar]
  71. Yamamoto, T.; Mizuao, K.; Tamori, I.; Ogata, A.; Nifuku, M.; Michalska, M.; Prieto, G. Catalysis-assisted plasma technology for carbon tetrachloride destruction. IEEE Trans. Ind. Appl. 1996, 32, 100–105. [Google Scholar] [CrossRef]
  72. Liang, W.J.; Li, J.; Li, J.X.; Zhu, T.; Jin, Y.Q. Formaldehyde removal from gas streams by means of NaNO2 dielectric barrier discharge plasma. J. Hazard. Mater. 2010, 175, 1090–1095. [Google Scholar] [CrossRef] [PubMed]
  73. He, C.; Cao, L.; Liu, X.; Fu, W.; Zhao, J. Catalytic behavior and synergistic effect of nonthermal plasma and CuO/AC catalyst for benzene destruction. Int. J. Environ. Sci. Technol. 2015, 12, 3531–3540. [Google Scholar] [CrossRef]
  74. Ding, H.-X.; Zhu, A.-M.; Lu, F.-G.; Xu, Y.; Zhang, J.; Yang, X.-F. Low-temperature plasma-catalytic oxidation of formaldehyde in atmospheric pressure gas streams. J. Phys. D 2006, 39, 3603–3608. [Google Scholar] [CrossRef]
  75. Chang, C.L.; Bai, H.; Lu, S.J. Destruction of styrene in an air stream by packed dielectric barrier discharge reactors. Plasma Chem. Plasma Process. 2005, 25, 641–657. [Google Scholar] [CrossRef]
  76. Abd Allah, Z.; Sawtell, D.; Kasyutich, V.; Martin, P.A. FTIR and QCL diagnostics of the decomposition of volatile organic compounds in an atmospheric pressure dielectric packed bed plasma reactor. J. Phys. Conf. Ser. 2009, 157, 1–9. [Google Scholar] [CrossRef]
  77. Fitzsimmons, C.; Ismail, F.; Whitehead, J.C.; Wilman, J.J. The chemistry of dichloromethane destruction in atmospheric-pressure gas streams by a dielectric packed-bed plasma reactor. J. Phys. Chem. A 2000, 104, 6032–6038. [Google Scholar] [CrossRef]
  78. Okubo, M.; Yamamoto, T.; Kuroki, T.; Fukumoto, H. Electric air cleaner composed of nonthermal plasma reactor and electrostatic precipitator. IEEE Trans. Ind. Appl. 2001, 37, 1505–1511. [Google Scholar] [CrossRef]
  79. Lin, C.-H.; Bai, H. Energy-effectiveness of nonthermal plasma reactors for toluene vapor destruction. J. Environ. Eng. 2001, 127, 648–654. [Google Scholar] [CrossRef]
  80. Chang, C.L.; Lin, T.S. Decomposition of toluene and acetone in packed dielectric barrier discharge reactors. Plasma Chem. Plasma Process. 2005, 25, 227–243. [Google Scholar] [CrossRef]
  81. Li, D.; Zhang, D.; Wu, Y.; Li, J.; Li, G. A study of removing chlorobenzene by the synergistic effect of catalysts and dielectric-barrier discharge Driven by bipolar pulse-power. Plasma Sci. Technol. 2008, 10, 94–99. [Google Scholar]
  82. Okubo, M.; Kuroki, T.; Kametaka, H.; Yamamoto, T. Odor control using the ac barrier type plasma reactors. IEEE Trans. Ind. Appl. 2000, 868–875. [Google Scholar]
  83. Massines, F.; Gouda, G. A comparison of polypropylene-surface treatment by filamentary, homogeneous and glow discharges in helium at atmospheric pressure. J. Phys. D 1998, 31, 3411–3420. [Google Scholar] [CrossRef]
  84. Manley, T.C. The electric characteristics of the ozonator discharge. Trans. Electrochem. Soc. 1943, 84, 83–96. [Google Scholar] [CrossRef]
  85. Chirumamilla, V.R.; Hoeben, W.; Beckers, F.; Huiskamp, T.; Van Heesch, E.J.M.; Pemen, A.J.M. Experimental investigation on the effect of a microsecond pulse and a nanosecond pulse on NO removal using a pulsed DBD with catalytic materials. Plasma Chem. Plasma Process. 2016, 36, 487–510. [Google Scholar] [CrossRef]
  86. Ogata, A.; Shintani, N.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Decomposition of benzene using a nonthermal plasma reactor packed with ferroelectric pellets. IEEE Trans. Ind. Appl. 1999, 35, 753–759. [Google Scholar] [CrossRef]
  87. Tamaki, K.; Yoshida, H.; Katayama, T.; Kaido, C. Oxidation of nitrogen monoxide by corrona discharges. 1. Effect of discharge conditions. Nippon Kagaku Kaishi 1979, 11, 1582–1588. [Google Scholar] [CrossRef]
  88. Zheng, C.; Zhu, X.; Gao, X.; Liu, L.; Chang, Q.; Luo, Z.; Cen, K. Experimental study of acetone removal by packed-bed dielectric barrier discharge reactor. J. Ind. Eng. Chem. 2014, 20, 2761–2768. [Google Scholar] [CrossRef]
  89. Chang, J.S.; Kostov, K.G.; Urashima, K.; Yamamoto, T.; Okayasu, Y.; Kato, T.; Iwaizumi, T.; Yoshimura, K. Removal of NF3 from semiconductor process flue gases by tandem packed bed plasma-adsorbent hybrid systems. IEEE Trans. Ind. Appl. 1998, 36, 1845–1852. [Google Scholar]
  90. Takaki, K.; Urashima, K.; Chang, J.S. Ferro-electric pellet shape effect on C2F6 removal by a packed-bed-type nonthermal plasma reactor. IEEE Trans. Plasma Sci. 2004, 32, 2175–2183. [Google Scholar] [CrossRef]
  91. Jin, Q.; Jiang, B.; Han, J.; Yao, S. Hexane decomposition without particle emission using a novel dielectric barrier discharge reactor filled with porous dielectric balls. Chem. Eng. J. 2016, 286, 300–310. [Google Scholar] [CrossRef]
  92. Urashima, K.; Kostov, K.G.; Chang, J.S.; Okayasu, Y.; Iwaizumi, T.; Yoshimura, K.; Kato, T. Removal of C2F6 from a semiconductor process flue gas by a ferroelectric packed-bed barrier discharge reactor with an adsorber. IEEE Trans. Ind. Appl. 2001, 37, 1456–1463. [Google Scholar] [CrossRef]
  93. Tu, X.; Gallon, H.J.; Whitehead, J.C. Transition behavior of packed-bed dielectric barrier discharge in argon. IEEE Trans. Plasma Sci. 2011, 39, 2172–2173. [Google Scholar] [CrossRef]
  94. Holzer, F.; Kopinke, F.D.; Roland, U. Influence of ferroelectric materials and catalysts on the performance of non-thermal plasma (NTP) for the removal of air pollutants. Plasma Chem. Plasma Process. 2005, 25, 595–611. [Google Scholar] [CrossRef]
  95. Ogata, A.; Einaga, H.; Kabashima, H.; Futamura, S.; Kushiyama, S.; Kim, H.H. Effective combination of nonthermal plasma and catalysts for decomposition of benzene in air. Appl. Catal. B 2003, 46, 87–95. [Google Scholar] [CrossRef]
  96. Oda, T.; Takahahshi, T.; Yamaji, K. Nonthermal plasma processing for dilute VOCs decomposition. IEEE Trans. Ind. Appl. 2002, 38, 873–878. [Google Scholar] [CrossRef]
  97. Chang, J.S.; Chakrabarti, A.; Urashima, K.; Arai, M. The effects of barium titanate pellet shapes on the gas discharge characteristics of ferroelectric packed bed reactors. In Proceedings of the 1998 Annual Report Conference on Electrical Insulation and Dielectric Phenomena, Atlanta, GA, USA, 25–28 October 1998; Volume 2, pp. 485–488. [Google Scholar]
  98. Takaki, K.; Urashima, K.; Chang, J.S. Scale-up of ferro-electric packed bed reactor for C2F6 decomposition. Thin Solid Films 2006, 506, 414–417. [Google Scholar] [CrossRef]
  99. Bradford, M.C.J.; Vannice, M.A. Estimation of CO heats of adsorption on metal surfaces from vibrational spectra. Ind. Eng. Chem. Res. 1996, 35, 3171–3178. [Google Scholar] [CrossRef]
  100. Lee, B.Y.; Park, S.H.; Lee, S.C.; Kang, M.; Choung, S.J. Decomposition of benzene by using a discharge plasma-photocatalyst hybrid system. Catal. Today 2004, 93–95, 769–776. [Google Scholar] [CrossRef]
  101. Takahashi, A.; Yang, F.H.; Yang, R.T. Aromatics/aliphatics separation by adsorption: New sorbents for selective aromatics adsorption by π-complexation. Ind. Eng. Chem. Res. 2000, 39, 3856–3867. [Google Scholar] [CrossRef]
  102. Gandhi, M.S.; Ananth, A.; Mok, Y.S.; Song, J.I.; Park, K.H. Effect of porosity of α-alumina on non-thermal plasma decomposition of ethylene in a dielectric-packed bed reactor. Res. Chem. Intermed. 2014, 40, 1483–1493. [Google Scholar] [CrossRef]
  103. Holzer, F.; Roland, U.; Kopinke, F.D. Combination of non-thermal plasma and heterogeneous catalysis for oxidation of volatile organic compounds. Part 1. Accessibility of the intra-particle volume. Appl. Catal. B 2002, 38, 163–181. [Google Scholar] [CrossRef]
  104. Ogata, A.; Ito, D.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Removal of dilute benzene using a zeolite-hybrid plasma reactor. IEEE Trans. Ind. Appl. 2001, 37, 959–964. [Google Scholar] [CrossRef]
  105. Zhu, R.; Mao, Y.; Jiang, L.; Chen, J. Performance of chlorobenzene removal in a nonthermal plasma catalysis reactor and evaluation of its byproducts. Chem. Eng. J. 2015, 279, 463–471. [Google Scholar] [CrossRef]
  106. Dey, G.R.; Sharma, A.; Pushpa, K.K.; Das, T.N. Variable products in dielectric-barrier discharge assisted benzene oxidation. J. Hazard. Mater. 2010, 178, 693–698. [Google Scholar] [CrossRef] [PubMed]
  107. Gandhi, M.S.; Ananth, A.; Mok, Y.S.; Song, J.I.; Park, K.H. Time dependence of ethylene decomposition and byproducts formation in a continuous flow dielectric-packed plasma reactor. Chemosphere 2013, 91, 685–691. [Google Scholar] [CrossRef] [PubMed]
  108. Ogata, A.; Shintani, N.; Yamanouchi, K.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Effect of water vapor on benzene decomposition using a nonthermal-discharge plasma reactor. Plasma Chem. Plasma Process. 2000, 20, 453–467. [Google Scholar] [CrossRef]
  109. Zhao, D.-Z.; Li, X.-S.; Shi, C.; Fan, H.-Y.; Zhu, A.-M. Low-concentration formaldehyde removal from air using a cycled storage-discharge (CSD) plasma catalytic process. Chem. Eng. Sci. 2011, 66, 3922–3929. [Google Scholar] [CrossRef]
  110. Huang, H.; Ye, D.; Leung, D.Y.C. Abatement of toluene in the plasma-driven catalysis: Mechanism and reaction kinetics. IEEE Trans. Plasma Sci. 2011, 39, 877–882. [Google Scholar] [CrossRef]
  111. Ogata, A.; Yamanouchi, K.; Mizuno, K.; Kushiyama, S.; Yamamoto, T. Decomposition of benzene using alumina-hybrid and catalyst-hybrid plasma reactors. IEEE Trans. Ind. Appl. 1999, 35, 1289–1295. [Google Scholar] [CrossRef]
  112. Trinh, H.Q.; Mok, Y.S. Plasma-catalytic oxidation of acetone in annular porous monolithic ceramic-supported catalysts. Chem. Eng. J. 2014, 251, 199–206. [Google Scholar] [CrossRef]
  113. Horvath, G.; Mason, N.J.; Polachova, L.; Zahoran, M.; Moravsky, L.; Matejcik, S. Packed bed DBD discharge experiments in admixtures of N2 and CH4. Plasma Chem. Plasma Process. 2010, 30, 565–577. [Google Scholar] [CrossRef]
  114. Futamura, S.; Zhang, A.; Einaga, H.; Kabashima, H. Involvement of catalyst materials in nonthermal plasma chemical processing of hazardous air pollutants. Catal. Today 2002, 72, 259–265. [Google Scholar] [CrossRef]
  115. Futamura, S.; Zhang, A.; Yamamoto, T. Mechanisms for formation of inorganic byproducts in plasma chemical processing of hazardous air pollutants. IEEE Trans. Ind. Appl. 1999, 35, 760–766. [Google Scholar] [CrossRef]
  116. Lu, M.; Huang, R.; Wu, J.; Fu, M.; Chen, L.; Ye, D. On the performance and mechanisms of toluene removal by FeOx/SBA-15-assisted non-thermal plasma at atmospheric pressure and room temperature. Catal. Today 2015, 242, 274–286. [Google Scholar] [CrossRef]
  117. Zhu, X.; Gao, X.; Yu, X.; Zheng, C.; Tu, X. Catalyst screening for acetone removal in a single-stage plasma-catalysis system. Catal. Today 2014, 256, 108–114. [Google Scholar] [CrossRef]
  118. Zhu, X.; Tu, X.; Mei, D.; Zheng, C.; Zhou, J.; Gao, X.; Luo, Z.; Ni, M.; Cen, K. Investigation of hybrid plasma-catalytic removal of acetone over CuO/γ-Al2O3 catalysts using response surface method. Chemosphere 2016, 155, 9–17. [Google Scholar] [CrossRef] [PubMed]
  119. Oda, T.; Takahashi, T.; Yamaji, K. TCE decomposition by the nonthermal plasma process concerning ozone effect. IEEE Trans. Ind. Appl. 2004, 40, 1249–1256. [Google Scholar] [CrossRef]
  120. Harling, A.M.; Demidyuk, V.; Fischer, S.J.; Whitehead, J.C. Plasma-catalysis destruction of aromatics for environmental clean-up: Effect of temperature and configuration. Appl. Catal. B 2008, 82, 180–189. [Google Scholar] [CrossRef]
  121. Oda, T.; Yamaji, K.; Takahashi, T. Decomposition of dilute trichloroethylene by nonthermal plasma processing—Gas flow rate, catalyst, and ozone effect. IEEE Trans. Ind. Appl. 2004, 40, 430–436. [Google Scholar] [CrossRef]
  122. Jiang, L.; Nie, G.; Zhu, R.; Wang, J.; Chen, J.; Mao, Y.; Cheng, Z.; Anderson, W.A. Efficient degradation of chlorobenzene in a non-thermal plasma catalytic reactor supported on CeO2/HZSM-5 catalysts. J. Environ. Sci. 2016, 1–8. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, L.; Huang, B.; Su, Y.; Zhou, G.; Wang, K.; Luo, H.; Ye, D. Manganese oxides supported on multi-walled carbon nanotubes for selective catalytic reduction of NO with NH3: Catalytic activity and characterization. Chem. Eng. J. 2012, 192, 232–241. [Google Scholar] [CrossRef]
  124. Li, H.; Lu, G.; Dai, Q.; Wang, Y.; Guo, Y.; Guo, Y. Efficient low-temperature catalytic combustion of trichloroethylene over flower-like mesoporous Mn-doped CeO2 microspheres. Appl. Catal. B 2011, 102, 475–483. [Google Scholar] [CrossRef]
  125. Ogata, A.; Kim, H.H.; Futamura, S.; Kushiyama, S.; Mizuno, K. Effects of catalysts and additives on fluorocarbon removal with surface discharge plasma. Appl. Catal. B 2004, 53, 175–180. [Google Scholar] [CrossRef]
  126. Kim, H.H.; Ogata, A.; Schiorlin, M.; Marotta, E.; Paradisi, C. Oxygen isotope (18O2) evidence on the role of oxygen in the plasma-driven catalysis of VOC oxidation. Catal. Lett. 2011, 141, 277–282. [Google Scholar] [CrossRef]
  127. Kang, M.; Kim, B.J.; Cho, S.M.; Chung, C.H.; Kim, B.W.; Han, G.Y.; Yoon, K.J. Decomposition of toluene using an atmospheric pressure plasma/TiO2 catalytic system. J. Mol. Catal. A 2002, 180, 125–132. [Google Scholar] [CrossRef]
  128. Zhu, T.; Li, J.; Liang, W.; Jin, Y. Synergistic effect of catalyst for oxidation removal of toluene. J. Hazard. Mater. 2009, 165, 1258–1260. [Google Scholar] [CrossRef] [PubMed]
  129. Bahri, M.; Haghighat, F.; Rohani, S.; Kazemian, H. Impact of design parameters on the performance of non-thermal plasma air purification system. Chem. Eng. J. 2016, 302, 204–212. [Google Scholar] [CrossRef]
  130. Fan, X.; Zhu, T.; Wan, Y.; Yan, X. Effects of humidity on the plasma-catalytic removal of low-concentration BTX in air. J. Hazard. Mater. 2010, 180, 616–621. [Google Scholar] [CrossRef] [PubMed]
  131. Ran, L.; Wang, Z.; Wang, X. The effect of Ce on catalytic decomposition of chlorinated methane over RuOx catalysts. Appl. Catal. A 2014, 470, 442–450. [Google Scholar] [CrossRef]
  132. Lee, H.; Lee, D.H.; Song, Y.H.; Choi, W.C.; Park, Y.K.; Kim, D.H. Synergistic effect of non-thermal plasma-catalysis hybrid system on methane complete oxidation over Pd-based catalysts. Chem. Eng. J. 2015, 259, 761–770. [Google Scholar] [CrossRef]
  133. Chen, L.; Zhang, X.; Huang, L.; Lei, L. Application of in-plasma catalysis and post-plasma catalysis for methane partial oxidation to methanol over a Fe2O3-CuO/γ-Al2O3 catalyst. J. Nat. Gas Chem. 2010, 19, 628–637. [Google Scholar] [CrossRef]
  134. Aerts, R.; Tu, X.; Van Gaens, W.; Whitehead, J.C.; Bogaerts, A. Gas purification by nonthermal plasma: A case study of ethylene. Environ. Sci. Technol. 2013, 47, 6478–6485. [Google Scholar] [PubMed]
  135. Burg, S.P.; Burg, E.A. Role of ethylene in fruit ripening. Plant Physiol. 1962, 37, 179–189. [Google Scholar] [CrossRef] [PubMed]
  136. Harling, A.M.; Glover, D.J.; Whitehead, J.C.; Zhang, K. Novel method for enhancing the destruction of environmental pollutants by the combination of multiple plasma discharges. Environ. Sci. Technol. 2008, 42, 4546–4550. [Google Scholar] [CrossRef] [PubMed]
  137. Li, J.; Ma, C.; Xu, X.; Yu, J.; Hao, Z.; Qiao, S. Efficient elimination of trace ethylene over nano-gold catalyst under ambient conditions. Environ. Sci. Technol. 2008, 42, 8947–8951. [Google Scholar] [CrossRef] [PubMed]
  138. Jiang, N.; Li, J.; Shang, K.; Lu, N.; Wu, Y. Enhanced degradation of benzene in surface/packed-bed hybrid discharge system: Optimization of the reactor structure and electrical parameters. IEEE Trans. Plasma Sci. 2016, 44, 657–664. [Google Scholar] [CrossRef]
  139. Guerra, G.; Lemma, A.; Lerda, D.; Martines, C.; Salvi, G.; Tamponi, M. Benzene emissions from motor vehicle traffic in the urban area of milan: Hypothesis of health impact assessment. Atmos. Environ. 1995, 29, 3559–3569. [Google Scholar] [CrossRef]
  140. Jones, A.P. Indoor air quality and health. Atmos. Environ. 1999, 33, 4535–4564. [Google Scholar] [CrossRef]
  141. Coates, J.D.; Chakraborty, R.; Lack, J.G.; O’Connor, S.M.; Cole, K.A.; Bender, K.S.; Achenbach, L.A. Anaerobic benzene oxidation coupled to nitrate reduction in pure culture by two strains of Dechloromonas. Nature 2001, 411, 1039–1043. [Google Scholar] [CrossRef] [PubMed]
  142. Futamura, S.; Gurusamy, A. Synergy of nonthermal plasma and catalysts in the decomposition of fluorinated hydrocarbons. J. Electrostat. 2005, 63, 949–954. [Google Scholar] [CrossRef]
  143. Huang, H.B.; Ye, D.Q.; Fu, M.L.; Feng, F. Da Contribution of UV light to the decomposition of toluene in dielectric barrier discharge plasma/photocatalysis system. Plasma Chem. Plasma Process. 2007, 27, 577–588. [Google Scholar] [CrossRef]
  144. Zhu, T.; Wan, Y.D.; Li, J.; He, X.W.; Xu, D.Y.; Shu, X.Q.; Liang, W.J.; Jin, Y.Q. Volatile organic compounds decomposition using nonthermal plasma coupled with a combination of catalysts. Int. J. Environ. Sci. Technol. 2011, 8, 621–630. [Google Scholar] [CrossRef]
  145. Gibbs, B.F.; Mulligan, C.N. Styrene toxicity: An ecotoxicological assessment. Ecotoxicol. Environ. Saf. 1997, 38, 181–194. [Google Scholar] [CrossRef] [PubMed]
  146. Kuroki, T.; Hirai, K.; Kawabata, R.; Okubo, M.; Yamamoto, T. Decomposition of adsorbed xylene on adsorbents using nonthermal plasma with gas circulation. IEEE Trans. Ind. Appl. 2010, 46, 672–679. [Google Scholar] [CrossRef]
  147. Carpenter, C.P.; Geary, D.L.; Myers, R.C.; Nachreiner, D.J.; Sullivan, L.J.; King, J.M. Petroleum hydrocarbon toxicity studies XVII. Animal response to n-nonane vapor. Toxicol. Appl. Pharmacol. 1978, 44, 53–61. [Google Scholar] [CrossRef]
  148. Nilsen, O.G.; Haugen, O.A.; Zahlsen, K.; Halgunset, J.; Helseth, A.; Aarset, H.; Eide, I. Toxicity of n-C9 to n-C13 alkanes in the rat on short term inhalation. Pharmacol. Toxicol. 1988, 62, 259–266. [Google Scholar] [CrossRef] [PubMed]
  149. Niaz, K.; Bahadar, H.; Maqbool, F.; Abdollahi, M. A review of environmental and occupational exposure to xylene and its health concerns. EXCLI J. 2015, 14, 1167–1186. [Google Scholar] [PubMed]
  150. Saulich, K.; Müller, S. Removal of formaldehyde by adsorption and plasma treatment of mineral adsorbent. J. Phys. D 2013, 46, 45201. [Google Scholar] [CrossRef]
  151. US Environmental protection agency (EPA) (Office if Air and Radiation). Introduction to Indoor Quality, a Reference Manual; EPA: Washington, DC, USA, 1991; Volume EPA/400/3.
  152. National Research Council of the National Academics (US). Review of the Environmental Protection Agency’s Draft IRIS Assessment of Formaldehyde; National Research Council of the National Academics: Washington, DC, USA, 2011. [Google Scholar]
  153. Barker, J.R.; Herstrom, A.A.; Tingey, D.T. Formaldehyde: Environmental partitioning and vegetation exposed. Water Air Soil Pollut. 1996, 86, 71–91. [Google Scholar] [CrossRef]
  154. Flowers, L.; Broder, M.W.; Forsyth, C. Toxicological Review of Acetone; Environmental Protection Agency: Washington, DC, USA, 2003.
  155. Urashima, K.; Chang, J.S. Removal of volatile organic compounds from air streams and industrial flue gases by non-thermal plasma technology. IEEE Trans. Dielectr. Electr. Insul. 2000, 7, 602–614. [Google Scholar] [CrossRef]
  156. Zhang, A.; Futamura, S.; Yamamoto, T. Nonthermal plasma chemical processing of bromomethane. J. Air Waste Manag. Assoc. 1999, 49, 1442–1448. [Google Scholar] [CrossRef] [PubMed]
  157. Langård, S.; Rognum, T.; Fløtterød, Ø.; Skaug, V. Fatal accident resulting from methyl bromide poisoning after fumigation of a neighbouring house leakage through sewage pipes. J. Appl. Toxicol. 1996, 16, 445–448. [Google Scholar] [CrossRef]
  158. Bolt, H.M.; Gansewendt, B. Mechanisms of carcinogenicity of methyl halides. Crit. Rev. Toxicol. 1993, 23, 237–253. [Google Scholar] [CrossRef] [PubMed]
  159. Ames, B.N.; Gold, L.S. Animal cancer tests and cancer prevention. J. Natl. Cancer Inst. Monogr. 1992, 125–132. [Google Scholar]
  160. Basu, S. Carbon tetrachloride-induced lipid peroxidation: Eicosanoid formation and their regulation by antioxidant nutrients. Toxicology 2003, 189, 113–127. [Google Scholar] [CrossRef]
  161. Howards, P. Handbook of Environmental Fate and Exposure Data for Organic Chemicals; Lewis Publishers: Chelsea, MI, USA, 1989; Volume 2. [Google Scholar]
  162. Harling, A.M.; Wallis, A.E.; Whitehead, J.C. The effect of temperature on the removal of DCM using non-thermal, atmospheric-pressure plasma-assisted catalysis. Plasma Process. Polym. 2007, 4, 463–470. [Google Scholar] [CrossRef]
  163. Abd Allah, Z.; Whitehead, J.C.; Martin, P. Remediation of dichloromethane (CH2Cl2) using non-thermal, atmospheric pressure plasma generated in a packed-bed reactor. Environ. Sci. Technol. 2014, 48, 558–565. [Google Scholar] [CrossRef] [PubMed]
  164. Bell, B.P.; Franks, P.; Hildreth, N.; Melius, J. Methylene chloride exposure and birthweight in Monroe County, New York. Environ. Res. 1991, 55, 31–39. [Google Scholar] [CrossRef]
  165. Abedi, K.; Ghorbani-Shahna, F.; Jaleh, B.; Bahrami, A.; Yarahmadi, R.; Haddadi, R.; Gandomi, M. Decomposition of chlorinated volatile organic compounds (CVOCs) using NTP coupled with TiO2/GAC, ZnO/GAC, and TiO2-ZnO/GAC in a plasma-assisted catalysis system. J. Electrostat. 2015, 73, 80–88. [Google Scholar] [CrossRef]
  166. Karuppiah, J.; Reddy, P.M.K.; Reddy, L.E.; Subrahmanyam, C. Catalytic non-thermal plasma reactor for decomposition of dilute chlorobenzene. Plasma Process. Polym. 2013, 10, 1074–1080. [Google Scholar] [CrossRef]
  167. Wang, C.; Xi, J.Y.; Hu, H.Y.; Yao, Y. Advantages of combined UV photodegradation and biofiltration processes to treat gaseous chlorobenzene. J. Hazard. Mater. 2009, 171, 1120–1125. [Google Scholar] [CrossRef] [PubMed]
  168. Ogata, A.; Mizuno, K. Methane decomposition in a barium titanate packed-bed nonthermal plasma reactor. Plasma Chem. Plasma Process. 1998, 18, 363–373. [Google Scholar] [CrossRef]
  169. Futamura, S.; Zhang, A.; Prieto, G.; Yamamoto, T. Factors and intermediates governing byproduct distribution for decomposition of butane in nonthermal plasma. IEEE Trans. Ind. Appl. 1998, 34, 967–974. [Google Scholar] [CrossRef]
  170. Zhang, H.; Li, K.; Sun, T.; Jia, J.; Lou, Z.; Feng, L. Removal of styrene using dielectric barrier discharge plasmas combined with sol-gel prepared TiO2 coated γ-Al2O3. Chem. Eng. J. 2014, 241, 92–102. [Google Scholar] [CrossRef]
  171. Zhang, L.; Yang, D.; Wang, W.; Wang, S.; Yuan, H.; Zhao, Z.; Sang, C.; Jia, L. Needle-array to plate DBD plasma using sine AC and nanosecond pulse excitations for purpose of improving indoor air quality. Sci. Rep. 2016, 6, 25242. [Google Scholar] [CrossRef] [PubMed]
  172. Wallis, A.E.; Whitehead, J.C.; Zhang, K. The removal of dichloromethane from atmospheric pressure nitrogen gas streams using plasma-assisted catalysis. Appl. Catal. B 2007, 74, 111–116. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of packed bed non-thermal plasma reactor and its cross-sectional view (a) Power source; (b) Packing material in discharge gap; (c) Dielectric barrier; (d) High voltage (or powered electrode) and (e) Ground electrode.
Figure 1. Schematic diagram of packed bed non-thermal plasma reactor and its cross-sectional view (a) Power source; (b) Packing material in discharge gap; (c) Dielectric barrier; (d) High voltage (or powered electrode) and (e) Ground electrode.
Catalysts 07 00113 g001
Figure 2. V-Q Lissajous figure in the (a) absence and (b) presence of ceramic Raschig rings (RR), voltage and current waveform in the (c) absence and (d) presence of ceramic Raschig rings (RR) at applied voltage of 12 kV (50 Hz and sine wave). Reprinted from [63]. Copyright (2016), with permission from Elsevier.
Figure 2. V-Q Lissajous figure in the (a) absence and (b) presence of ceramic Raschig rings (RR), voltage and current waveform in the (c) absence and (d) presence of ceramic Raschig rings (RR) at applied voltage of 12 kV (50 Hz and sine wave). Reprinted from [63]. Copyright (2016), with permission from Elsevier.
Catalysts 07 00113 g002
Figure 3. Effect of presence of packing material on the power to the plasma reactor (frequency of power supply = 60 Hz; packing material: BaTiO3 pellets). Reprinted from [78]. Copyright (2016), with permission from IEEE.
Figure 3. Effect of presence of packing material on the power to the plasma reactor (frequency of power supply = 60 Hz; packing material: BaTiO3 pellets). Reprinted from [78]. Copyright (2016), with permission from IEEE.
Catalysts 07 00113 g003
Figure 4. Waveforms of applied voltage and discharge current for different dielectric constant of spherical pellets (a) 104; (b) 5000 and (c) 650 (applied voltage = 14 kV; frequency = 60 Hz). Reprinted from [90]. Copyright (2016), with permission from IEEE.
Figure 4. Waveforms of applied voltage and discharge current for different dielectric constant of spherical pellets (a) 104; (b) 5000 and (c) 650 (applied voltage = 14 kV; frequency = 60 Hz). Reprinted from [90]. Copyright (2016), with permission from IEEE.
Catalysts 07 00113 g004
Figure 5. Effect of dielectric constant on the removal efficiency of benzene (Initial benzene concentration: 200 ppm; carrier gas: dry air; flow rate: 0.2 L/min; shape and size of the packing material: sphere of φ2 mm, numbers in legend box refer to dielectric constant value, M = Mg2TiO4, C = CaTiO3, S = SrTiO3, B = BaTiO3). Reprinted from [86]. Copyright (2016), with permission from IEEE.
Figure 5. Effect of dielectric constant on the removal efficiency of benzene (Initial benzene concentration: 200 ppm; carrier gas: dry air; flow rate: 0.2 L/min; shape and size of the packing material: sphere of φ2 mm, numbers in legend box refer to dielectric constant value, M = Mg2TiO4, C = CaTiO3, S = SrTiO3, B = BaTiO3). Reprinted from [86]. Copyright (2016), with permission from IEEE.
Catalysts 07 00113 g005
Figure 6. Waveforms of applied voltage and discharge current for different pellet shape such as (a) sphere; (b) cylinder and (c) hollow cylinder (packing material = BaTiO3; dielectric constant = 104; applied voltage = 12 kV; frequency = 60 Hz). Reprinted from [90]. Copyright (2016), with permission from IEEE.
Figure 6. Waveforms of applied voltage and discharge current for different pellet shape such as (a) sphere; (b) cylinder and (c) hollow cylinder (packing material = BaTiO3; dielectric constant = 104; applied voltage = 12 kV; frequency = 60 Hz). Reprinted from [90]. Copyright (2016), with permission from IEEE.
Catalysts 07 00113 g006
Figure 7. Pathways of toluene degradation in the gas phase and on the FeOx/SBA-15 surface. Reprinted from [116]. Copyright (2017), with permission from Elsevier.
Figure 7. Pathways of toluene degradation in the gas phase and on the FeOx/SBA-15 surface. Reprinted from [116]. Copyright (2017), with permission from Elsevier.
Catalysts 07 00113 g007
Table 1. Source of different VOCs and their effect on human and environment.
Table 1. Source of different VOCs and their effect on human and environment.
CompoundSourcesHealth DisorderEnvironmental IssuesRef.
MethaneUnburned natural gasBreathing problemGreenhouse gas[132,133]
ButaneFuel, fuel additive, aerosol propellant, refrigerant, chemical feedstockDrowsiness, euphoria, fluctuation in blood pressure, memory loss, irregular heartbeatAffects aquatic ecosystem[57]
Ethylene (Ethene)Engine exhaust, petrochemical derivative, thermal power plant, food industry, production facility of polymers and other chemicalsAnesthetic illnessPhotochemical smog, ground level ozone[102,134,135,136,137]
HexaneSolvent in industry, gasoline vehiclesNausea, vertigo, bronchial and intestinal infection, affects central nervous system and can be fatalHaze[91]
BenzenePetrochemical waste gas, unleaded gasoline, production of ethylbenzene and styreneDrowsiness, dizziness, unconsciousness, anemia, leukemia, blood diseases, Carcinogen, affects central nervous system and cause respiratory disorderSmog, affects aquatic ecosystem, contaminates water and soil[48,49,69,73,138,139,140,141]
Toluene (Methyl benzene)As a solvent for production of paints, paint thinners, nailpolish, medicines, dyes, explosives, detergents, spot removers, lacquers, adhesives, rubber, antifreeze, printing, leather tanning, pharmaceutical industry, feedstock in chemical processesTiredness, confusion, weakness, memory loss, nausea, loss of appetite, hearing loss, loss of color vision, sleepiness, unconsciousness and deathStratospheric ozone depletion, climatic changes, affects the potable and agricultural water[38,127,128,142,143,144]
Styrene (Vinyl benzene)Petrification material for resin, plastic and medicineMemory loss, difficulties in concentration and learning, brain and liver damage, mutagen and potential carcinogenAffects aquatic organisms and ground water, forms ground level ozone[75,145]
Xylene (dimethyl benzene)Solvent and fixative in pathology laboratoriesEye, nose and throat irritation, difficulty in breathing, impaired lung function, central nervous system impairment and adversely affects the nervous systemPollutes the ground water[146,147,148,149]
FormaldehydeProduction facilities of urea-formaldehyde, phenol-formaldehyde resins and some building materials such as plywood, chip board and paneling, adhesives, resins, preservatives, cosmetics, incomplete combustion in power plants and automobiles, incineratorsIrritation of eyes and respiratory tracts, nausea, headache, fatigue, dullness, thirst and suspected carcinogenDetrimental to vegetation[42,72,74,150,151,152,153]
AcetaldehydeTobacco smoke, production facilities of acetic acid, pyridine and pyridine base, peracetic acid, pentaerythritol, butylene glycol and chloralIrritation of eye, nose and throat, causes conjunctivitis, coughing, central nervous system depression, eye and skin burns, dermatitis and delayed pulmonary edemaPhotochemical smog[60,78]
Acetone (2-propanone)Chemical production processes, vehicle emission, tobacco smoke, incineration of waste materials and as a solvent in cleaning and paintingDizziness, unconsciousness, nauseaOxygen depletion in aquatic system[88,117,118,154]
PerfluoroethaneEtching process and cleaning of process chambers in semiconductor industryAffects respiratory systemAffects vegetation growth and fresh water lakes, greenhouse effect[90,92,98,155]
Bromomethane (Methyl bromide)Fumigant on agricultural fields and quarantine facilitiesHeadache, nausea, vomiting, dizziness, blurred vision, impairment of coordination, twitching, permanent debilitation or death, carcinogen (excessive exposure)Depletes stratospheric ozone layer[156,157,158,159]
Carbon tetrachlorideSolvent in industrial chemicalsHepatic injuryDepletes stratospheric ozone layer[71,160,161]
DichloromethaneSolvent in various industrial applicationsHeadache, nausea, dullness, dizziness, pulmonary irritation, affects the central nervous system; abortion, affect the birth weight, carcinogen (excessive exposure)Stratospheric ozone depletion, climatic changes[38,162,163,164]
ChlorobenzeneProduction of phenol and dichloro-diphenyl-trichloroethane (DDT), high boiling solvents, precursor for 2-nitrophenol synthesis, solvent in pharmaceutical and dye industriesNervous and kidney damage, dizziness, neurasthenia, increase the risk of cancerPrecursor of secondary aerosols, photochemical smog[19,81,105,122,165,166,167]
TrichloroethyleneWashing agent to remove oil and hydrophobic contaminantsCarcinogenic, nerve paralyzing, sick house syndromeDepletion of ozone[37,65,121]
Table 2. Overview of published papers on abatement of VOCs using packed bed non-thermal plasma reactors.
Table 2. Overview of published papers on abatement of VOCs using packed bed non-thermal plasma reactors.
CompoundPacking MaterialSize/ShapeDielectric ConstantCarrier GasFlow Rate 1 (L/min)Initial Conc. 2 (ppm)Removal Efficiency (%)Energy Density/FrequencyBy-ProductsRef.
Methane (CH4)BaTiO3φ1 mm/sphere5000N2110006610 kJ/L (50 Hz)--[156]
BaTiO3φ1 mm/sphere15,000Air0.210001711 WNOx, N2O, NH3, HCN, C2H2[168]
N23411 W (50 Hz)
No packing----2.5% + O2 + He200 cm3/min250032165 J/L--[132]
2 wt % Pd/SiO2φ0.45–0.6 mm/sphere3.935169 J/L
2 wt % Pd/TiO2‘’ 38030172 J/L
2 wt % Pd/Al2O3‘’954163 J/L (4 kHz)
Butane (CH3CH2CH2CH3)No packing----N2 + O2 (5%)1200065.56100 W (400 Hz)N2O, CH3CHO[57]
GW----90.37
TiO2/GWφ3 mm/sphere--90.31
α-Al2O3φ3 mm/sphere--83.97
ZrO2φ3 mm/sphere--84.83
GO/GW 4----85.60
BaTiO3φ1 mm/sphere--Dry N20.51000948 kV (50 Hz)CH4, C2H6, CH2=CH2, CH3CHOC3H8, CH2=CHCH3, Butenes, CH3(CH2)2CHO, CH3CCH2CH3, Ethyloxirane, 2,3-dimethyloxirane, CH3COCH3, CH3CHO, Methyl ether, Butanols, CH3CN, CH3NO2, NO, N2O, NO2[169]
Humid N2 (2%)89
Dry air90
Humid air (2%)88
Ethylene (CH2=CH2)No packing----Air11898781819 J/L--[107]
α-Al2O3--91001013 J/LN2O
SiO2--3.91001054 J/LN2O
ZrO2--25991932 J/LCH4, C2H2, HCHO, N2O
GW--3.41001900 J/L (400 Hz)CH4, C2H2, HCHO, N2O
Non-porous α-Al2O3--9.45% O2 + N2118989980 WN2O, HCHO, CH3CHO[102]
Porous α-Al2O3--9.410045 W--
Porous zeolite----10045 W (400 Hz)N2O, polymer like deposit
PropyleneGlass beadsφ5 mm/sphere--Dry air10150795.5 W (60 Hz)--[59]
2060
HexaneNo packing----N2 + 15% O20.5293–36764399 J/L2-hexanone, 3-hexanone[91]
Quartzφ1.79–2.29 mm/sphere3.959415 J/L
Al2O3‘’--67422 J/L
Porous Al2O3‘’9.970389 J/L
Benzene (C6H6)2 wt % Ni/TiO2φ1.8 mm/sphere--Air (100 °C)4–1020093210 J/LNO2, N2O, Ozone[41]
2 wt % Ag/TiO2φ1.8 mm/sphere--86215 J/L
0.5 wt % Ag/γ-Al2O3φ2 mm/sphere--83240 J/L
0.5 wt % Pt/γ-Al2O3φ3.2 × 3.6 mm/cylinder--95210 J/L
0.5 wt % Pd/γ-Al2O3φ3.2 × 3.6 mm/cylinder--78240 J/L
2 wt % Ag/H-Yφ1.5 × ~5 mm/cylinder--76200 J/L
Ferrieriteφ1.5 × ~5 mm/cylinder--79200 J/L (500 Hz)
BaTiO3φ5 mm/sphere----1500>5130 W--[45]
BaTiO3 pellets--4000Dry air0.2200602 W (50 Hz)C2H2[46]
Humid air42
N240
Humid N223
BaTiO3φ2 mm/sphere10,000Dry air0.2–320052.9386 J/LHCOOH[47]
Humid air (4500 ppm)37.1384 J/L
Ag/TiO2φ2 mm/sphere--Dry air89.1383 J/L
Humid air (4500 ppm)86391 J/L
No packing----Dry air1.61300 mg/m35412 kV/cmOzone[48]
TiO2/RRφ5 mm/sphere 10 mm (length) 1 mm wall Thickness--8013.6 kV/cm
3% Ag/TiO2φ1.4–1.7 mm/sphere--O20.5520010010 J/L (50 Hz)Ozone[49]
γ-Al2O3φ2–3 mm/sphere--Air332080800 J/L (8 kHz)NO2, NO[50]
Pt/γ-Al2O3φ2–3 mm/sphere--89
1 wt % Ag/TiO2φ1.8 mm/sphere100 °CDry air T = 100 °C4203–21092.5236 J/LHCOOH, NO2, N2O[52]
RR 5φ5 × 10 mm/cylinder“--Dry air14 mm/s600 mg/m37810 kV/cmOzone, Aromatic polymers[53]
TiO2/RR1500 mg/m38914 kV/cm
600 mg/m39810 kV/cm
1500 mg/m39814 kV/cm (60 Hz)
Glassφ 2 mm/sphere--Dry air0.54004910 W--[54]
BaTiO3φ2 mm/sphere10,000Dry air2203–21064400 J/L--[55]
TiO2φ1.8 mm/sphere--81390 J/L
1 wt % Ag/TiO2φ1.8 mm/sphere--88380 J/L
1 wt % Pt/TiO2φ1.8 mm/sphere--78390 J/L (100 Hz–1.5 kHz)
Glassφ3 mm/sphere--Dry air--2403425 kVhydroquinone, heptanoic acid, 4-nitrocatechol, phenol, 4-phenoxy-phenol[56]
BaTiO3φ2 mm/sphere4000Humid air (0.5%)0.220096.85.2 W (50 Hz)C2H2, Ozone, N2O, HNO3[66]
No packing----Air0.510034350 J/L--[67]
MnO2φ1.18–1.7 mm/sphere--‘’41360 J/L
TiO2φ1.7 mm/sphere--0.145300 J/L (50 Hz)
1 wt % Ag/TiO2φ1.8 mm/sphere--Dry air420081.5197 J/LNO2, N2O, HCOOH[69]
N294.5247 J/L
Ar-O29694 J/L
CuO/AC 6----Humid air (50%)18,857 h−124090.670 J/LNOx[73]
AC----70.5
MnO/AC---->85
MnO2/AC---->85
BaTiO3φ2 mm/sphere15,000Dry air0.220098.88 WNOx, N2O, C2H2, Ozone[86]
SrTiO3φ2 mm/sphere260099.78 W
CaTiO3φ2 mm/sphere20065.36.7 W
Mg2TiO4φ2 mm/sphere2067.6 W (50 Hz)
Al2O3/BaTiO3φ2 mm/φ1 mm/sphere10/5000Humid air (0.5%)0.22001003 kJ/L--[95]
BaTiO3φ2 mm/sphere5000~1002.6 kJ/L
CaTiO3φ2 mm/sphere200702 kJ/L
No packing----Air0.210040140 J/L (900 Hz)C6H5OH[100]
TiO2/Glass beadsφ5 mm/sphere--50
BaTiO3φ2 mm/sphere4000Humid air (0.5%)0.2200746 W (50 Hz)N2O, NOx, C2H2[44]
BaTiO3/Al2O3φ2 mm/φ1 mm/sphere87
BaTiO3φ1 mm/sphere4000Humid air (0.5%)0.2200806 W (50 Hz)N2O[104]
MS-3A/BaTiO3φ2 mm/φ1 mm/sphere72
MS-4A/BaTiO3‘’74
MS-5A/BaTiO3‘’85
MS-13X/BaTiO3‘’100
BaTiO3φ2 mm/sphere4000Humid air (0.5%)0.2200796 W (50 Hz)N2O, NOx, C2H2[111]
Al2O3φ1 mm/sphere--88.5
BaTiO3φ2 mm/sphere4000Dry air0.2200979 WN2O, NO2, C2H2[108]
Humid air (1%)839 W
N2938 W
Humid N2 (1%)7710 W (50 Hz)
No packing----Air0.540052340 J/L--[138]
BaTiO3φ5 × 5 mm/cylinder--73.5498 J/L
Glassφ5 mm/sphere--70.8468 J/L
--67400 J/L (50 Hz)
No packing----Dry air0.410003942 kV (50 Hz)--[119]
MnO2----58
Toluene (C6H5CH3)BaTiO3 pelletφ1 mm/sphere5000Dry air0.86010014 kVOzone[38]
0.2571008 kV (60 Hz)
0.5 wt % Pt/γ-Al2O3----Air4–1015098210 J/L (500 Hz)--[41]
2 wt % Ag/TiO2----92
BaTiO3 beads--4000Dry air0.2200662 W (50 Hz)CH4, C6H6, C6H5CH3, C2H2[46]
Humid air59
N245
Humid N237
Glass beads----Dry air0.31524036172 J/L (110 Hz)C6H5COOH[58]
Glass beadsφ5 mm--Air5200305.5 W--[59]
‘’5011 W
257.55.5 W
‘’7611 W (60 Hz)
RRφ9.2 mm × 10.5 mm/cylinder--Humid air (30%)8.5--93687 J/LOzone[61]
NaNO2/RR--95786 J/L
BaTiO3/RR--98818 J/L 25 kHz)
No packing----Air0.0680060--CH2O2, C2H4O2, C6H6 C6H5CH2OH, C6H5COOH, 5-methyl-2-nitrophenol (as deposit)[62]
OMS-2----70
10 wt % Cu/OMS-2----80
Al2O3----75
1 wt % Au/Al2O3----90
Nb2O5----96
No packing----Air10700 mg/m336280 J/L--[43]
Glass----45330 J/L
Ceramic RR----58350 J/L
Al2O3----72380 J/L
BaTiO3/Al2O3----59360 J/L
TiO2/Al2O3----52390 J/L
BaTiO3/TiO2/Al2O3----71340 J/L (50–500 Hz)
No packing----Air0.331200 mg/m372660 J/LOzone[63]
Ceramic RRφ5 mm/cylinder--97450 J/L (50 Hz)
No packing----Air0.16489 J/L 50 Hz)C6HCH2OH, C6H5CHO, C6H5COOH[68]
NiO/γ-Al2O3----88.8
MnO2/γ-Al2O3----78.8
CeO2/γ-Al2O3----77.1
Fe2O3/γ-Al2O3----73.2
CuO/γ-Al2O3----70.7
γ-Al2O3----63.5
No packing----Humid air (95%)15006611.6 kV (60 Hz)N2O[79]
Glass pelletsφ5 mm591
No packing----Air0.611003511 W--[80]
Glass pelletsφ5 mm--789 W
TiO2/glass pellets--7710 W (60 Hz)
No packing----Air0.250741527 J/LOzone[110]
TiO2/Al2O3/Ni foam----94
Glass beads----O2410004013 kV (60 MHz)--[127]
3 wt % TiO2/glass beads----70
TiO2/γ-Al2O3----80
γ-Al2O35–7 mm/sphere--Air21000 mg/m3750.7 kJ/L (150 Hz)Ozone[128]
TiO2/γ-Al2O3--88
5 wt % MnO2/γ-Al2O3--85
10 wt % MnO2/γ-Al2O3--98
15 wt % MnO2/γ-Al2O3--98
No packing----Air0.1200504 W (12,5 kV)--[103]
γ-Al2O321553
No packing----N20.310043192 J/L (50 Hz)--[116]
N2 + 3% O251
Air27
SBA-15----N250
N2 + 3% O286
Air92
N261
N2 + 3% O298
3% FeOx/SBA-15----Air97
N250
SiO2----Air45
Styrene (C6H5CH=CH2)No packing----Dry air0.6408914 kV (60 Hz)C6H6,Toluene, Benzene acetaldehyde[75]
Glassφ5 mm--50
Al2O3‘’--70
Pt-Pd/Al2O3‘’--96
No packing----Dry air0.51000 mg/m35412.5 kV (10 kHz)NO, NO2[170]
γ-Al2O3--70
P25/γ-Al2O3--89
TiO2/γ-Al2O3--97
Xylene ((CH3)2C6H4)BaTiO3φ5 mm----14909528 W--[45]
BaTiO3 pellets--4000Dry air0.2200762 W (50 Hz)CH4, C6H6, C6H5CH3, C2H2[46]
Humid air71
N249
Humid N249
Formaldehyde (HCHO)No packing----Humid air (1%) T = 70 °C0.60514077108 J/L (50 Hz)NO, NO2, N2O, HCOOH[42]
α-Al2O3--84
γ-Al2O3--95
No packing---Humid air (30%)8.550 mg/m370670 J/LOzone[72]
3000 ppm NaNO2/RRφ9.2 × 10.5 mm/cylinder--75678 J/L
5000 ppm NaNO2/RR‘’--80664 J/L
8000 ppm NaNO2/RR‘’--86656 J/L (15 kHz)
No packing----Humid air (1%)T = 70 °C0.60527699108 J/L (50 Hz)NO, N2O, NO2[74]
Ag/CeO2 pellets----99
γ-Al2O3----78
Fused silica----57
No packing----Dry air157.772645 J/LHCOOH[18]
CeO2----66615 J/L
CuO----44575 J/L
Cu1Ce3----91583 J/L
Cu1Ce1----94485 J/L
Cu3Ce1----‘’87530 J/L (10 kHz)
No packing----Dry air0.21029340 W--[171]
γ-Al2O39942 W (9 kHz)
Acetaldehyde (CH3CHO)γ-Al2O3φ5 mm--5% O2, N21100010046 W (60 Hz)N2O, NO2, CH3NO2, CH3NO3, HCN, CH3OC2O5 CH3COOH CH3OH[60]
BaTiO3 pelletsφ1.7–2 mm10,000Humid air (4%)11009216 kVOzone, NO (low), NO2 (low), N2O (low)[82]
29618 kV
Dry N219714 kV
28816 kV (60 Hz)
No pellet----Dry air1100875.8 WNO (low), NO2 (low), N2O (low)[78]
BaTiO3 pelletsφ1.7–2 mm10,000Dry air922.4 W
Humid air (2%)792.4 W
Acetone (CH3COCH3)No packing----Dry air0.210040.83441 J/LNO2, N2O, Ozone[88]
Glass pelletsφ1 mm3.954.01534 J/L
α-Al2O3--9.559566 J/L
γ-Al2O3--12.680640 J/L (50 Hz)
γ-Al2O3----Dry air1118464.11287 J/LHCHO, HCOOH[117]
10 wt % CuOx/γ-Al2O3----94.31012 J/L
10 wt % CeOx/γ-Al2O3----761036 J/L
10 wt % CoOx/γ-Al2O3----891047 J/L
10 wt % MnOx/γ-Al2O3----891051 J/L
10 wt % NiOx/γ-Al2O3----791051 J/L (10 kHz)
γ-Al2O3--12.6Dry air--500 mg/m35825 W (10.2 kHz)HCOOH, HCHO, NO2, N2O[118]
2.5 wt % CuO/γ-Al2O3--65
5 wt % CuO/γ-Al2O3--68
7.5 wt % CuO/γ-Al2O3--64
10 wt % CuO/γ-Al2O3--58
No packing----Dry air0.611001519 W--[80]
Glassφ5 mm4516 W
TiO2/glass4715 W (60 Hz)
Perfluoroethane (CF3-CF3)BaTiO3φ3.3 mm/sphere10,000N20.03300071 W (60 Hz)--[90]
φ3.2 mm × 4 mm/cylinder12
BaTiO3φ3.2 mm × 4 mm/cylinder10,000N2115007514 kVCF4, CHF3[98]
30007715 kV (60 Hz)
BaTiO3φ3 mm/sphere10,000Dry N20,02530004512 kV (60 Hz)CF4, NO2, N2O, SiF4[92]
Humid N2 (70%)13
Bromomethane (CH3Br)BaTiO3φ1 mm5000N21100010011 kJ/LCH4, CH2Br2, BrCN[156]
Humid N2 (2%)8814 kJ/L
Air7413 kJ/L
Humid air (2%)7014 kJ/L (50 Hz)
Carbon tetrachloride (CCl4)No pellet----Dry air0.2450–56047.14.24 kV--[71]
BaTiO3 pelletsφ1 mm--1004.24 kV
Ni/BaTiO3--1004,1 kV
SrTiO3 pellets--93–1003.76 kV (18 kHz)
Dichloromethane (CH2Cl2)BaTiO3φ1.4–2.8 mm--N21500541000 J/L (100Hz)HCN, HCl, CCl4, N2O, COCl2 NOCl, COCl2, HCN, HCl, CCl4, COCl2, HCl, COCl2[163]
Air55
N2 + 3% O274
Ar70
Ar + 21% O290
BaTiO3φ5 mm----14608928 W--[45]
BaTiO3--2000–10,000N2 + 3% O2150075900 J/L (=15 W) (20 kHz)N2O, NO2, NO, ClNO[76]
BaTiO3φ3.5 mm--N215001866 J/L (= 1.1 W) (10 kHz)HCN, Cl2, HCl, COCl2, N2O, NO2, HCl, Cl2[77]
Air12
BaTiO3φ3.5 mm--Air1500271.1 WHOCl, NO, NO2, N2O[172]
γ-Al2O3/BaTiO3--360.9 W (10.25–13.25 kHz)
Chlorobenzene (C6H5Cl)No packing----Dry air0.361250 mg/m3847 kJ/L (10 kHz)--[122]
CeO2/HZSM-5φ0.3–0.5 mm--96
No packing----Dry air0.361250 mg/m3325 kV (10 kHz)Benzene derivatives, nitrogenous organics, small molecule alkyds[105]
CuO/MnO2φ0.3–0.5 mm--70
CeO2/HZSM-5----72
Ag/TiO2----53
No packing----Dry air0.7425331280 J/L--[81]
MS-4Aφ2 mm--42529960 J/L
6 wt % MnO2/ALP----41527960 J/L (50 Hz)
Trichloroethylene (Cl2C=CHCl)No packing----Dry air0.41000930.9 W--[65]
V2O5/TiO2φ5 mm/sphere--980.3 W
Cu-ZSM-5φ3–6 mm × 1 mm thick/disc--960.1 W (50 Hz)
No packing----Dry air0.41000830.4 WOzone[121]
TiO2φ2–3 mm/sphere--910.6 W
V2O5/TiO2‘’--940.6 W
WO3/TiO2‘’--950.5 W (50 Hz)
No packing----Dry air0.4100971.4 WCH3COOH, C2H2Cl3NO[96]
TiO2φ0.5–1 mm--981 W
TiO2φ2–3 mm--960.6 W
TiO2φ4–5 mm × 1 mm (disk)--980.4 W (50 Hz)
No packing----Dry air0.41000900.5 WCl2HC-COCl[119]
MnO2----950.6 W (50 Hz)--
BaTiO3φ1 mm/sphere--Dry air11000865.1 kV/cmOzone, NO, NO2[70]
Humid air (2%)705.2 kV/cm
N2943.9 kV/cm
1 The unit of flow rate is L/min (unless and otherwise specified); 2 the unit of initial conc. is ppm (unless and otherwise specified); 3 “ = same as previous line; 4 GO, graphene oxide, GW, glass wool; 5 RR, Raschig rings; 6 AC, activated carbon.

Share and Cite

MDPI and ACS Style

Veerapandian, S.K.P.; Leys, C.; De Geyter, N.; Morent, R. Abatement of VOCs Using Packed Bed Non-Thermal Plasma Reactors: A Review. Catalysts 2017, 7, 113. https://doi.org/10.3390/catal7040113

AMA Style

Veerapandian SKP, Leys C, De Geyter N, Morent R. Abatement of VOCs Using Packed Bed Non-Thermal Plasma Reactors: A Review. Catalysts. 2017; 7(4):113. https://doi.org/10.3390/catal7040113

Chicago/Turabian Style

Veerapandian, Savita K. P., Christophe Leys, Nathalie De Geyter, and Rino Morent. 2017. "Abatement of VOCs Using Packed Bed Non-Thermal Plasma Reactors: A Review" Catalysts 7, no. 4: 113. https://doi.org/10.3390/catal7040113

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop