Vous êtes sur la page 1sur 30

654

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7o. 5, October 2000

Air Pollution Control by Electrical Discharges


R. Hackam and H. Akiyama
Department of Electrical and Computer Engineering The University of Kumamoto Kumamoto, Japan

ABSTRACT
Air pollution caused by gas emission of pollutants produced from a wide range of sources including coal, oil and gas burning power plants, diesel engines, paper mills, steel and chemical production plants must be reduced drastically and urgently, as mandated by recent worldwide national legislation which recently are being reinforced increasingly by international agreements. Non-thermal plasmas in which the mean energy of the electrons is substantially higher than that of the gas offer advantages in reducing the energy required to remove the pollutants. The electrical energy supplied into the discharge is used preferentially to create energetic electrons which are then used to produce radicals by dissociation and ionization of the carrier gas in which the pollutants are present. These radicals are used to decompose the pollutants. There are two technologically promising techniques for generating non-thermal plasmas in atmospheric gas pressure containing the pollutants, namely electron beam irradiation and electrical discharge techniques. Both techniques are undergoing intensive and continuous development worldwide. This is done to reduce the energy requirement for pollutant removal, and therefore the associated cost, as well as to obtain a better understanding of the physical and chemical processes involved in reducing the pollutants. In the present paper only electrical discharge techniques are reviewed and emphasis is given to the more recent published work. The paper summarizes the chemical reactions responsible for the removal of the major polluting constituents of NO, NO2 and SO1 encountered in flue gases and exhaust emissions. The constructional features of the various types of electrical discharge reactors commonly employed in the removal of gas pollutants as well as pilot systems used in industrial plants are described briefly. The results on the removal efficiency of the various pollutants including hydrocarbons and volatile compounds and their dependency on the type of discharge reactor, the type and the magnitude of the applied voltage (dc, ac and pulsed), the polarity of the voltage (dc and pulsed), the effect of the pulse width, the initial concentration of the pollutants, the addition of ammonia, argon and other hydrocarbons, the gas flow rate, the residence time of the pollutants in the discharge reactor, the gas temperature and on the type of the gas constituents will be reviewed. The removal of pollutants using arc plasmas will be discussed. The specific energy density which is supplied into various forms of electrical discharges to reduce the pollutants will be discussed. The energy required to remove the pollutants is expected to be one of the main considerations in selecting the technology to be used to remove the pollutants and therefore it is of prime importance.

1 INTRODUCTION
IR pollution caused by gas emission of pollutants produced by a variety of sources including coal, oil and gas burning electric power generating Plants, motor vehicles, diesel engine exhausts, paper mills and steel and chemical production plants [1-111 must be substantially reduced as mandated by recent national legislation and interna-

tional agreements. These pollutants are major causes of acid rain, planet warming, smog and also increasingly are being thought of as detrimental to human health, to vegetation growth and to fresh water lakes. Several techniques have been applied in recent years to remove pollutants from the air with various degrees of success. Non-thermal plasmas in which the mean energy of the electrons is substantially higher than that of the ions and the neutrals offer considerable advantage in

1070-9878101 $3.00 02000 IEEE

IEEE Transactionson Dielectrics and Electrical Insulation


reducing the energy requirements to remove the pollutants. Under certain conditions the electrical energy fed into the discharge is used preferentially to create energetic electrons instead of heating the ions and the neutral gas molecules. The energetic electrons are employed to dissociate and ionize the carrier gas molecules in which the pollutants are present to produce radicals. The radicals are then used to decompose the pollutants. There are essentially two main techniques that have been developed for creating non-thermal plasmas at atmospheric pressure containing the pollutants. These methods are electron beam irradiation and electrical discharge. Both methods are undergoing continuous development at an intensive level worldwide due to the urgency of achieving mandated legislated emission reduction. The current efforts to improve the technologies of pollution reduction have the main objectives of reducing the energy requirements as well as achieving a better understanding of the collision processes involved in the removal of the toxic pollutants in the flue gases and in simulated gas mixtures containing pollutants. Electron beam irradiation techniques have been investigated widely and were applied in pilot projects [2,10-201. However, in the present paper only electrical discharge techniques will be reviewed and particular emphasis will be given to the more recent published work. Only reported results, available in the public domain, are in this review. A reference to the electron beam irradiation method will be made for comparison purpose wherever applicable. This paper will begin with a brief review of the main reactions which are thought to be responsible for the removal of the major polluting constituents of NO, NO2 and SO2 encountered in flue gases and vehicle exhausts. A brief description of laboratory and pilot industrial systems will be given. The essential constructional features of the various types of electrical discharge reactors commonly employed in the removal of gas pollutants in flue gases will be discussed. The dependence of the removal efficiencies of the pollutants on the geometry of the reactors, the type of the discharge reactor (without dielectric barrier, with dielectric barrier and surface discharges), the type and the magnitude of the applied voltage (dc, ac and pulsed), the gas flow rate, the residence time of the pollutants in the discharge reactor, the gas temperature, the type of gas constituents which carry the pollutants, the polarity of the applied voltage (dc and pulsed), the effects of the pulse width, the initial concentration of the pollutants, the addition of hydrocarbons, argon, ammonia and other compounds will be reviewed. The removal of pollutants using arc plasmas will be discussed. The most recently reported results on the specific energy consumption to remove NO, NO2 and SO2 and the prevalent toxic pollutants also will be reviewed.

Vol. 7 N o . 5, October 2000

655

usually given in parts per million (ppm) of the pollutant of the main carrier gas. 1ppm in l.01x105Pa of the carrier gas such as air or other mixtures at 273 K is equal to a density of 2 . 6 9 ~ 1 0 ~ ' mol/m3. The main carrier gas is usually air with enhanced amounts of water vapor (HzO), carbon dioxide (CO?),carbon monoxide (CO) and varying amounts of NO, and SO?. NO2 The application of HV to a pair of electrodes immersed in a gas containiig the pollutants creates high energy electrons which produce ions and free radicals. The main objective of producing the plasma is to provide the radicals N, 0 and OH which interact with the pollutant molecules [21]. Depending on the constituents of the gas mixture the following reactions have been suggested widely to occur, in which the radicals N, 0, and OH are produced in the presence of water vapor 1221 etN2+ etNtN (1) e t Nz + 2 e t N t t N (2) e t 02 + e o ( ~ P )o ( ~ P ) (3) e t 0 --* 2 e t 0 + 2e + 0 + O+ 2 e t H20 i e t H t OH
e tH20 +2etOHtH'

+ + e + O(3P) + O('D)

(4)

(5) (6) (7)


"

'

"

" ' ;" .' '

"

'

'

"

''1

10 2

50

100

150

200

250

300

E (kV/cm)
Figure 1. Initial yield of neutrals in an electrical discharge from electron impact processes. Gas mixture: 7% N2, 5% 0 2 , 10% H20, and 15%CO,. Electric field corresponds to gas pressure of 1 . 0 1 ~ 1 0 ~ Pa and 300 K 1221.

2 CHEMICAL REACTIONS FOR REMOVALS OF POLLUTANTS


A substantially large number of studies have been directed at reducing or eliminating particularly the main polluting constituents in flue gases which are the NO, and SO,. The NO, is generally referred to as the sum of the concentrations of NO and NO2 while the SO, is The primarily composed of SO?. concentration of the pollutants are

Figure 1shows the initial yields of neutral species calculated by Penetrante [22] which are produced in an electrical discharge by electron impact processes in a gas mixture of 70% N2, 5% 02, 10% HzO and 15%CO2 at a pressure of 1 . 0 1 ~ 1 Pa~and 300 K. The values shown in 0 Figure 1which were derived from the excitation, dissociation and ion2 ization reactions of WZ, 0 and C02, were considered reliable but those from the dissociation of HzO were uncertain by a factor of < 2 [22]. Radicals may also be produced by other reactions such as electron attachment to and dissociation of water molecules as shown in Figure 2

656

Hackam et al.: Air Pollution Control by Electrical Discharges

10'1

, ,

'

"

' ' I

" "

'

' '

'

"

"

. '

'

'

' ' '

OH Production i

. mnuibubw tfam anachment

50

io0

150

200

250

300

E (kV/cm)
Figure 2. Calculated production of OH radicals in an electrical discharge by collision reactions with H20of electron attachment, excited oxygen atoms, dissociation by electrons and positive ions as a function of electric field. Gas mixture and other conditions are as in Figure 1
P21.

[221

e t H2O ---t H t OH 0 (1D) t H20 + 2 0 H And from reaction (5) by charge transfer processes [22] 0 t Nz.+ 0 t N2+ ' Ott 0 2 --t 0 t 02+

(8) (9)
(10) (11)

Figure 2 shows the production of the radical OH as a function of the electric field at 1 . 0 1 ~ 1 0 ~ in an electrical discharge from colliPa sion reactions with H20 of electron attachment, excited oxygen atoms, dissociation by electrons and positive ions [22]. It will be shown later in this paper that the addition of water to the gas mixture aids considerably the removal of NO, and SO,. This is partly due to the added production of the radical OH which then interacts with and reduces the pollutants. The radical OH may remove NO according to the following reaction [21-231 (12) where M is a third molecule which can be, say N2. In the latter case 0 ~ [5,23-261. NO2 may be the reaction rate Klz = 6 . 7 ~ 1 cm6/s ~ ~ reduced by the radical OH [2, 22,26,28,29] OH t NO2 t Nz
+ HN03 t

OH t NO t M

+ HN02 t

Reaction (15) is followed with the production of the radical H02 [2], HS03 t 0 2 + HOz t so3 (15) ~ ' with a rate coefficient of K15 = 4 ~ 1 0 cm3/s~[34]. And [2] so3 t H20 + (16) HO2 t NO + NO2 t OH (17) NO2 reacts with H02 to produce HN03 and the radical 0 [25] NO2 t H 0 2 t N 2 + HNOJ t N2 t 0 (18) with a reaction rate of Kla = ~ . I X ~ O - ~ [25]. cm6/s ~ Reactions (14) to (17) oxidize SO2 and NO which generate OH that also reacts with NO2 according to (13). NO is also removed by oxidation with the radical 0 [2,22,23,31,32] O t N O t M + N02tM (19) with a reaction rate of Klg = 6.9x1OW3' cm6/s [23]. The latter coefficient is in agreement with the value of 6 . 3 ~ 1 cm6/s ~ ~ 0 ~ reported by [301. Similarly [24,25,31,34], O t S O z t M + S03tM (20) with a reaction rate K20 = 8 . 2 ~ 1 0 - ~ ' cm6/s [24,25,33,34]. And [25] so3 t H2O + H2S04 (21) And [6] SO2 t H20 + H2S03 (22) H2SO3 is easily oxidized to H2SO4 [19]. Ammonia is added to neutralize H2S04 to form ammonium sulfate [6,21,25], HzS04 t 2 NH3 + (NH4)2S04 (23) NO may be reduced directly by the reaction with N [15,21,22,28,32], NtNO+ N2tO (24) cm3/s [28]. with a reaction rate of K24 = 5.9~10-l1 NO2 which initially is present in the flue gas as well as that produced in reaction (19), may decompose into NO by [23,28,31], N t N O z + 2N0 (25) with a reaction rate K25 = 9.0x10-12 cm3/s [23,28]. In the presence of ammonia (NH3) the following reactions have been suggested [23,28] NH3 t NO2 + HN02 t NH2 (26) with a reaction rate of KZS 5 . 2 ~ 1 cm3/s ~ ~ = 0 ~ [23,28]. N is also removed by [23-25,27,28], NH2 t NO + H20 t N2 (27) with a reaction rate of K27 = 2 . 1 ~ 1 0 cm3/s [23-25,27,28]. NH2 ~~' IS created by (26) and NH3 t OH + NH2 t H20 (28) with a reaction rate K Z S= 3 ~ 1 0 - cm3/s [30]. l~

(13) The development of electrical discharge technology for the reduc-~~ where K13 = 2 . 6 ~ 1 0 cm6/s [5,25-281. And for SO2 [2, 21,30,34, tion of pollutants was spurred in part by the high capital cost of the 351 equipment associated with electron beam irradiation systems. The elecOH t SO2 + HS03 (14) trical discharge techniques involve a creation of a plasma in a flowing gas within a reactor. There are different designs of discharge reactors with a reaction rate of K14 = 7 . 4 ~ 1 0 cm3/s [34,35]. ~~' which attempt to produce very high electric field by employing thin wires, sharp needles and narrow pipes. The HV electrodes are placed

N2

3 ELECTRICAL DISCHARGE REACTORS

IEEE Transactions on Dielectrics and Electrical Insulation

between grounded plates which can be planar, in the form of a rect- and typically 15 to 200 ns duration is applied. The low voltangular duct, or cylindrical, either metal or dielectric wrapped with a age electrode could be cylindrical (Figure 3(a)) or a flat plate and is metal foil. However, cylindrical reactors have been used very widely in grounded. simulation studies because of the relative simplicity of their construcThe short duration pulses produce energetic electrons which genertion. ate radicals that decompose the pollutants. Due to the limited duration of the narrow pulses, the discharge is short lived and consequently the gas +. a .? --+ gas ion and gas molecules remain effectively at the ambient temperature. in out Thus only a negligibly small amount of the supplied electrical energy is dissipated to heat the gas. This type of reactor (Figure 3(a)) can be high-voltage adapted readily to industrial precipitators which use wire-plate geometry [10,11]. (I ', The second type of reactor is shown schematicallyin Figure 3@)[36] wire anode dielectric tube and is called 'dielectric barrier discharge' reactor or 'silent discharge' reactor. This reactor employs a thin dielectric sheet of glass, quartz or \ I ceramic, either planar or cylindricalin shape, of 1to 5 mm in thick-w.2+ gas ness and on its outer surface a metal foil which is grounded. The central s 7 ; out HV electrode usually is a thin wire. A high alternating voltage, 50 or 60 Hz or higher frequency of 1to 100 kHz is usually employed with hlgh-voltage this reactor [lo, 1 1but also a pulsed voltage has been used extensively 1 [36]. Dielectric barrier discharge reactors have been used widely in the metal sheet around dielectric tube commercial production of ozone for water purification for a long time wire anode [37,38]. In this type of reactor (Figure 3(b)) microdischarge streamers dielectric pellets are generated in the gap which self-extinguish when the electric field \ is reduced by the buildup of surface charge on the dielectric. When a short pulsed voltage is used with this type of reactor, a higher electric field to pressure ratio ( E / P )can be achieved which results in a larger production of radicals to destroy the pollutants. This leads to a more efficient operation of this dielectric barrier discharge reactor. dielectric tube The third type of discharge reactors is the dielectric pellet-bed dis(C ) metal sheet wire anode charge reactor shown in Figure 3(c) [36]. It uses high permittivity diaround dielectric lube electric pellets which are held between two metal electrodes to which Figure 3. Three types of discharge reactors. (a) concentric coaxial elecusually HV ac is applied, but pulsed operations also have been emtrodes for pulsed corona, (b) coaxial or planar electrodes for dielectric ployed. In Figure 3(c) the voltage is applied between the central wire barrier (also called silent) discharge (dielectric barrier either on one or and the outer cylinder, but in other arrangements two mesh forming both electrodes when plates are employed) and (c) cylindrical pelletbed reactor (electrodes may be in the form of flat metal mesh placed flat electrodes are placed on both sides of the cylinder, holding the pelon both sides of the cylinder containing the pellets) [36]. lets therein. The pellets are polarized and a high field is created at each pellet upon application of the external voltage and microdischargesare Embedded eleclrode (Surface H V. eleclrode) created at each pellet. Suitace eleclrode
N N

-L+
/
\

Vol. 7 N o . 5, October 2000

657

- -

',\

Ceotial eleclfode (Sllenl H V eleclrode)

Figure 4. Schematic diagram of combined surface discharge and dielectric barrier discharge reactor (281. There are essentially four types of discharge reactors three of which are shown in Figure 3 [36] and the fourth in Figure 4 [28,23]. The construction of the Dulse corona reactor (Figure 3(a)) uses a thin wire to which a very short, positive or negativek' pulse of 10 to 60 kV

The fourth type of discharge reactor is the 'surface discharge' reactor. It is composed of a series of strip electrodes placed on the surface of alumina ceramic, while a thin metal electrode is embedded inside the ceramic. The ceramic could be in a planar [39] or a cylindrical form 1281. HV ac is applied to generate the surface discharge, but pulsed operations were used also. Figure 4 shows a schematic diagram of a cylindrical surface discharge reactor where strip electrodes are placed on the surface, an electrode is embedded inside a ceramic and a central rod electrode [28]. A surface discharge is formed between the strip electrodes; in this arrangement a dielectricbarrier (or silent) discharge is also formed between the central and the surface electrodes [28].
A schematic arrangement of a positive pulsed corona discharge reactor to remove the NO,, and SO2 from the flue " Eases of an iron ore sintering plant followed by an electrostaticprecipitator (ESP) is shown in Figure 5 [6]. The ESP uses negative HV dc to collect the products

- -

658
FLUE CAS FROM

Hackam et al.: Air Pollution Control by Electrical Discharges


Plate elecimle
Exhausr
---, - - -I

H.V. (PositiveDC)
4-

Gasnow

Figure 5. A schematic diagram of a pilot system for removal of NOx, SO2 and solid particulates from the flue gases of iron-ore sintering plant [6].

u u
Addiriod 8 s

UULJ

Figure 7. Schematic diagram of it discharge reactor employing a pipe as the HV electrode. Ammonia and other gases are fed to the gas stream via the pipe electrode [40,41]
O.S.C.

Plrmnsl rompuar

Figure 8. Schematicdiagram of a laboratoryapparatus for the study of pollutant removal using pulsed corona in a coaxial discharge reactor with a concentric wire [32]. Figure 6. Positive pulsed discharge reactor for removal of NO, and SO1 from the flue gases of an iron-ore sintering plant followed with an electrostatic precipitator to collect ammonium sulfate, ammonium nitrate and other solid particulates. Wire diameter, 3 mm. Up to 5000 Nm3/h of the flue gas can be treated N, normalized to l.01x105Pa and 0C [6]. ders. The percentage of added moisture can be varied and is determined by adjusting the temperature of the water. The composition of the gas is analyzed at the exhaust of the reactor. The flow rate of the gas is adjustable and is measured using flow meters. The temperature of the gas can be raised by placing the reactor in an oven to simulate the actual temperatures of the flue gases in boilers and diesel exhausts.

A more effective coaxial discharge reactor for the removal of gas polof ammonium sulfate, ammonium nitrate and other solid particulates. lutants employs a spiral wire, usually wound on a hollow insulator rod. Both the discharge reactor and the ESP employ thin wires (3 mm in A typical discharge reactor that employs this principle of construction diameter) and plate electrodes combinations which are depicted in Fig- is shown in Figure 9 [42]. This reactor employs a copper wire, 1 mm ure 6 [6]. This reactor can treat up to 5000 Nm3/h (N-normalized to in diameter, wound on a polyvinyl chloride tube 3 cm in diameter and 0C and l.01x105Pa) of the flue gases. The ammonia and other addi- is placed concentric in a copper cylinder having 96 mm inner diameter tives are inserted into the gas stream at the inlet of the corona discharge and is 3.5 m long. reactor (Figure 5) [6]. Typical corona pulses of the voltage and current and the variation Another variation of the corona type discharge reactor which em- of the waveforms with pulse width are shown in Figures 10(a) and (b), ploys a pipe electrode and either positive or negative HV dc, is shown respectively [43]. The pulse width (or duration) is defined in the pulsed in Figure 7 [40,41]. In this reactor additives, such as ammonia, are power technology as the full-width-at-half-maximum (FWHM) of the injected into the gas stream via the pipe electrode (Figure 7). voltage pulse. It will be shown later in this paper that a short pulse is A typical laboratory setup for studying the removal of pollutants us- very desirable in obtaining a higher energy efficiency for the removal ing a concentric wire cylinder geometry is shown in Figure 8 [32]. The of pollutants. The electrical energy supplied into the discharge reactor simulated gas mixture is obtained from commercially available cylin- is calculated from the digitized signals of the applied voltage to and

IEEE Transactions on Dielectrics and Electrical InsulationI

Vol. 7 No. 5, October 2000

659
-40

O.S.C.

F
L

40

----- 100 ns

. -60

ns

ns

--80 ns

k
0
IC.

m m

>

F,
0

I ,

I,,

I I

I I I

100

, , I , ,, , I , , , , 200 300 Time, ns


,

I , ,

351111 mm

i/J
>

..... .._.

a
IC.

100

Figure 9. A laboratory setup employing a coaxial discharge reactor with spiral central electrode. Copper wire, 1 mm diameter, spiral pitch, 10 mm; central tube diameter, 30 nun; outer cylinder, 96 mm; reactor length, 3.5 m [42]. the discharge current in the reactor using J vidt, where v (in V) is the voltage, i (in A) the discharge current, and t is the time (ins). A typical dependence of the energy per pulse on the pulse width is shown in Figure 1 [43], where it can be seen that the energy pulse increases with 1 increasing pulse width. This is of importance when determining the energy density required for the removal of pollutants from flue gases. A typical magnetic pulse compressor (MPC), the type of which is increasingly being used to produce short pulse durations that are directly applied to discharge reactors for pollutant reductions is shown in Figure 12 [44].

s o
, ,-100, ,0, , , , ,E
IQ)

? ! L

t100

Time, ns

200

300

Figure 10. (a) Applied voltage to and (b) discharge current in a coaxial discharge reactor for different pulse widths. Discharge reactor: central wire, stainless steel 0.5 mm in diameter; outer cylinder, copper 76 mm in diameter; length of reactor 0.5 m. Gas mixture: 200 ppm NO, 5% Oz,4% HzO,91% Nz; gas flow rate 2.0 limin; pressure l.01x105Pa; temperature 25C [a].

__.._.___..... --- 5

DEVELOPMENT OF THE STREAMER DISCHARGE

It has been reported that applying short durations and fast rising positive pulses to a concentric coaxial wire-to-cylinderdischarge reactor resulted in superior room-temperature performance in the removal of NO and NO2 than using negative pulses [4]. Similarly in a discharge reactor which employed a needle or rod to plate electrodes geometries, positive pulsed corona performed better than negative corona in removing SO2 [7]. This has been attributed to the production of more uniform 0 100 200 300 Time, ns streamers that extended further into the gap for positive pulsed streamers [7]. For example it was shown that in a positive wire-planar geomFigure 11. Input energy to a coaxial discharge reactor per pulse for etry in l.01x105 Pa air, there were more streamer channels (8 chandifferent pulse widths. Curve 1,40 ns; curve 2, 60 ns; curve 3, 80 ns; nelsicm) compared to the negative wire (1 to 2 channels/cm) and the curve 4,100 ns; curve 5,120 ns. Gas mixture and reactor type are as in Figure 9 [U]. latter had a reduced branching compared to the positive pulsed corona [45]. For the same reason of more abundant and spread out streamers the positive pulsed corona led to a higher production of ozone than that in air after 3 pulses using a shorter duration of the applied voltage of 100 ns and a longer gap distance of 38 mm [47]. for the negative polarity in the wire-cylinder configuration [46]. Figure 13 shows the abundance of positive streamer discharges in Figure 15 shows the development of positive streamers in a flowing air when a single pulse of 60 kV and having 200 ns duration was applied nitrogen (2 l/min) filled coaxial discharge reactor (wire diameter 2 mm, to a wire of 1.5 mm in diameter and a plate with a gap of 20 mm [32]. inner diameter of outer cylinder 76 mm, length 500 mm) at a pressure The observed density of the streamers shown in Figure 13 was typically of 1 . 0 1 ~ 1 Pa, when successively a larger number of positive pulses 0~ 6 streamer channels/cm. Figure 14 shows the positive pulsed streamer (50 kV, 100 ns) were applied to the wire [48]. Figure 15 shows that the

660

Hackam et al.: Air Pollution Control by Electrical Discharges

Figure 12. Typical circuit diagram of an MPC used to produce short duration HV pulses for pollution control. GTO, gate turn off thyristor; C1, primary energy storage capacitor; SL1 and SL2 saturable inductors; PT, step-up pulse transformer; C3 peaking capacitor [44].

Figure 13. Positive pulsed streamer discharges in air between wire (top) and grounded plane electrode (bottom). Wire diameter 1.4 mm, gap distance 20 mm; applied voltage 60 kV; pulse duration 200 ns; pressure l.01x105 Pa; temperature 25T; single shot 1321.

Figure 15. Development of positive pulsed streamers in nitrogen in coaxial discharge reactor. Wire diameter 2 mm, inner diameter of the outer cylinder 76 mm, length of reactor 500 nun; applied voltage 50 kV;pulse duration 100ns;pressure 1 . 0 1 ~ 1 Pa; temperature 25C; 0~ gas flow rate 2 I/min; number of total applied pulses: a, 1;b, 3; c, 5; d, 7; e, 10; f, 30 1481. [4,7,25,40,41,49,50,51] or a combination of dc and pulsed corona discharges [7,26]. Generally dc corona discharges are less energy efficient than fast rising, short duration pulsed corona for the removal of pollutants from flue gases. Figure 16 shows a typical dc corona current us. applied positive dc voltage for the treatment of a flowing (5.3 l/min) mixture of gases containing 8016 N2,20% 0 2 and 155 ppm NO, with and without additions 2 of argon (Ar), ammonia (NH3) and Nz t 0 (401. The discharge reactor used has the electrode geometry as depicted in Figure 7. It will be observed from Figure 16 that the corona discharge current increased with increasing applied dc voltage, and at fixed voltage the current increased with increasing flow rate of the additives Ar and NH3 (Ar: NH3 = 973). The current increased further when additional N2 t 0 2 (N2: 0 2 = 8020) was injected into the simulated flue gas. Argon was employed to minimize the generation of negative ions in the corona discharge [5, 401. The increased corona current with increasing gas flow rate was attributed to stabilizing electrohydrodynamic flow effects [52]. At flow rates t0.4 l/min of the added NZ t 02, the corona discharge became unstable and was transformed into a spark discharge which was less effective and therefore not suitable for NO, removal. Figure 17 shows the corona discharge current in the flue gases from

Figure 14. Positive pulsed streamer discharges in air between wire (top) and grounded plane electrode (bottom). Wire diameter 1.0 mm, gap distance 38 nun; applied voltage 50 kV; pulse duration 100 ns; pressure l.01x105 Pa; temperature 25C; number of pulses 3 [47]. streamers gradually extend across the gap with increasing number of pulses from 1pulse (Figure 15(a)) to 30 pulses (Figure 15(f)).

DEVELOPMENT OF dc CORONA

HVDC to create corona discharges of either negative

There have been extensive studies of emission control employing [5,26,27], positive

IEEE Transactions on Dielectrics and Electrical InsulationI


Additional gas N2+02 Ar+NH3 0.8 I/min 01 \/min . A 0.8 Vmin 0 Vmin 0 0.4 l/nin 0 I/min

Vol. 7 No. 5, October 2000

661

increasing applied voltage. Further, the addition of a mixture of ammonia and argon (NH3: Ar = 3:97) led to a larger discharge current than that of adding a pure ammonia [5]. After 10 h of operation of this reactor the currents were observed to increase at the same appliedvoltage. This was attributed to the accumulation of particulate contaminants from the flue gases on the electrode surface leading to an enhanced electric field therein [5]. In this pollution control test which was carried out on the Fujisawa plant in Japan using a combination of electron beam irradiation and corona discharge the particulate of ammonium sulfate and ammonium nitrate were collected by an electrostatic precipitator and a baghouse filter. The cleaned flue gases were released through the stack by an induced draft fan [5].

a
Applied voltage (kV) Figure 16. Corona current vs. positive dc voltage in a simulated mixture of flue gases containing 80% N2,20% 02, and 155 ppm NO and with additives of argon and ammonia (NH,). Conditions: discharge reactor shown in Figure 7; distance between pipe electrode and duct wall 25 mm; concentration of NH3 in Ar 3%; flow rate of treated gas 5.3 l/min; gas pressure l.01x105 Pa; flow rates of additives are shown on the Figure [40].
c

A Needle

IoPlate

0
+

1 0
dc

20 30 bias voltage ( k V )

Figure 18. Corona discharge current us. applied positive dc voltage for dc, and for pulse t dc in rod-to-plate and needle-to-plategeometries. Needle tip radius of curvature 0.1 mm; rod diameter 5 mm with 2.5 mm radius of tip curvature; electrode spacing 4 cm; gas mixture air with 1666 ppm SO2 and 2.5% H20; pressure 1 . 0 1 ~ 1 Pa; temper0~ ature 22C; flow rate 1.2 I/ min; peak of pulse 45 kV; repetition rate
60 PPS [71
-10 -20 -30 -40 APPLIED VOLTAGE (kV)

-50

Figure 17. Corona discharge current os. negative applied voltage in flue gas from a heavy oil-fired boiler with additives of either ammonia or ammonia and argon. Conditions: Flue gas contained 200 to 1000 ppm SO2; 50 to 200 ppm NO,; 13 to 14% 02; gas temperature 65C; water vapor concentrationnot specified; flow rate 1200 m3/h. t 10 l/min of NH3; A 2 l/min of NH3 with 16 l/min of Ar [5]. a heavy oil-fired boiler vs. negative dc applied voltage with additives of either ammonia or ammonia and argon [5]. The flue gases contained concentrations of 200 to 1000 ppm of SO2 and 50 to 200 ppm of NO,. After applying water spray for cooling, the oxygen concentration was 13 to 14%. The temperature of the flue gases was reduced from an initial high value of 280 to 65C by using a heat exchanger and atomized water injection [5]. The additives of Ar and NH3 were injected into the corona discharge reactor via a hollow electrode (5 mm outside diameter and 10 cm long) which also formed the corona electrode. It will be observed from Figure 17 that the corona discharge increased with

For rod-to-plate and needle-to-plate discharge reactor geometries the dependence of the corona current on the applied dc voltage in a mixture of air, 1666 ppm of SO2 and 2.5% HzO, electrode gap disan tance of 4 cm, at a pressure of l.01x105Pa, 22C and a gas flow rate of 1.2 limin is shown in Figure 18 [7]. The corona current is shown for positive dc voltage and for a combined pulsed and dc voltages [7]. Figure 18 shows that under the same applied voltage the corona current was higher for the positive needle-to-plate than for the positive rod-toplate geometries. Under the same conditions the corona current was higher for the positive pulsed t positive dc voltages than for positive dc voltage (Figure 18) [7]. For positive dc cor0n.a using needle-to-plate electrodes at t30 kV, which was slightly below the sparking voltage and therefore the highest that could be used, 78% of SO2 was removed while for negative corona of even a higher voltage of -50 kV only 54% was removed. Thus positive dc voltage had a better performance than negative. Further, it was concluded that the removal efficiency of SO2 was poorer in the dc corona compared with the pulsed corona treatments [7].

662

Hackam et al.: Air Pollution Control by Electrical Discharges

REMOVAL OF POLLUTANTS USING dc VOLTAGE

The reduction of NO and NO2 in a simulated mixture of gases containing N2: 02: C02: NO = 83.96 8: 8: 0.04 as a function of applied negative dc voltage with ammonia (NH3) and methane (CHa) injections are shown in Figures 19 and 20, respectively [26]. The initial concentration of NO of 438 ppm decreased to 398 ppm when ammonia was injected in a ratio of NH3/NO = 0.5 without applying the voltage. Figure 19 shows that NO continued to decrease with increasing voltage. At the highest voltage of 34.7 kV, NO decreased to 197 ppm constituting a removal ratio of 55%. Using Fourier transform infrared (FTIR) measurements, trace amounts of HN03 and N20 were detected in agreement with other studies [9,24,27,53]. Figure 20 shows the concentrations of NO, NO2 and NO, as a function of the negative dc voltage with methane injection [26]. It will be observed that the initial NO concentration was not affected by the methane injection without the corona discharge. The maximum removal ratio of NO in the latter case was 52.7%which occurred at -30 kV (Figure 20) [26].

10

20

30

40

dc vokage [kq

400 500

r-7
INW
10

Figure 20. Concentration of NO,NO? and NO, as a function of negative dc voltage with methane (CHa) injection CH4/NO, = 1 Other conditions are as m Figure 19 [26]

E I

40

Voltage (kV)

O J

I
20
dc voltage [kv
30
40

Figure 19. Concentrations of NO,NO2 and NOx as a function of applied negative dc voltage with ammonia injection. Initial concentration of NO 438 ppm; NH3/NOx = 0.5; flow rate 5 I/min; gas mixture NZ: 02: C02: NO = 83.96 8: 8: 0.04; pressure l.01x105 Pa; discharge reactor plane to plane [26].

Figure 21. Dependence of the removal ratio of NO, on positive dc voltage with ammonia injection for clean and contaminated electrodes. Gas mixture: N2: 0 2 : CO2 = 89.4: 2.63: 7.9 t 200 ppm NO t 800 ppm SOz with 0.5% H20; pressure l.01x105 Pa; room temperature; gas flow rate 5 l/min; injected additives (NH3/(NO t SO2)) = 1.2; (NH3 (10%)t Ar) diluted with air [25].

In a simulated mixture of gases of N2: 02: CO2 = 89.4: 2.63 7.9 with 0.5% H20 containing 200 ppm NO and 800 ppm S 2 the deO pendence of the NO, and SO2 removal ratios on the applied positive dc voltage are shown with addition of ammonia in Figures 21 and 22, respectively [25]. The gas flow rate was 5 l/min and the controls of the pollutants were carried out at room temperature. It will be observed from Figure 21 that the NO, removal ratio increased with increasing applied positive dc voltage. 82% of NO, was removed at 24 kV The contaminated electrodes resulted in larger removal efficiency of NO, in agreement with other studies [54,55] and was attributed to surface reactions of aerosol particles with NH3 radicals and NO, [53]. Figure 22 shows the removal ratio of SO2 as a function of the applied positive dc voltage with the addition of ammonia at a concentration of

[NH3/(NO t SOz)] = 1.2 [25]. It is noteworthy, and as can be seen from Figure 22, that 85% of SO2 can be removed without applying any voltage. This was attributed to the reactions with SO2 of the aerosols generated by the adiabatic expansion cooling of air-Ar-NHs mixture which was injected into the simulated flue gas [53,56]. Figure 22 shows that the removal efficiency of SO2 increases slightly with increasing applied positive dc voltage to 23 kV, and then remains constant with further increase in the voltage when the electrodes are clean. The dominant reactions for SO2 removal are the formation of ammonium sulfate [(NH&S02] by particle surface reactions between NH3 and SO2 molecules [25,56]. After l h of operation of the discharge reactor, the electrode surfaces became contaminated with aerosol particle deposition and the removal efficiency of SO2 decreased (Figure 22) [25]. The dependence of the removal ratios of NO (a) and NO, (b) on the dc input power to the corona discharge is shown in Figure 23 for a sim-

<

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7 N o . 5, October 2000

663

G
-

8Q

+Clean electrode 0 Contaminatedelectrode

the power was calculated is shown in Fi ure 16 [40]. Fi re 23 shows the effect of the flow rate of the simulate(Hflue gases on t E removal ratio of NO (Figure 23(a)) and on NO,(Figure 23@)).It will be observed that both NO and NO, removal ratios decreased substantially with increasing the flow rate of the gas from 5 to 10 l/min. The gas residence time in the corona region was calculated to be 19 s for the higher flow rate of 10 l/min [40].
100

e! 0 BO "
U )

*.*..*.-*--.

s
B

75

Time (h)

Figure 24. Dependence of NOx reduction rate on time of application of positive dc voltage in a corona discharge reactor. Conditions: simulated flue gas: 80% NZt 20% 0 t 190 ppm NO at flow rate 4.9 l/min t 2 19%0 t 4.76%CO2 t 0.23% NH3 t 76% N?at flow rate of 0.84 l/min; 2 pressure l.01x105 Pa; discharge reactor as per Figure 7 [41].

I
0

(a)

8 12 Corana power (W)

16

I
0
@)

12

16

Corona power (W)

Figure 23. Dependence of removal ratios of NO (a) and NO,@) on dc input power to the corona discharge. Discharge reactor is shown in Figure 7. Gas mixture: 80% N2 t 20% 0 t 149 ppm NO at 5 and 2 10 l/min t 80% NZt 20% 0 at 0.8 I/min t Ar t NH3 (3%)at 0.1 I/min; 2 Pa; positive dc voltage; pressure 1 . 0 1 ~ 1 0 ~ 05 l/min, 10 l/min ~401. ulated flue gas mixture of N2: 0 = 80%: 20% t 149 ppm of NO at 5 and 2 10 l/min and using additional gases injected into the flue gas of 80% Nz t 20% 0 at 0.8 l/min and (97% Ar t 3% NH3) at 0.1 l/min [40]. The 2 discharge reactor used in this study [40] is shown in Figure 7 [40,41]. The dependence of the corona current on applied voltage from which

The dependence of the NO,removal ratio on the time of application of a positive dc voltage is shown in Figure 24 for a simulated mixture of gases containing 80% NZ t 20% 0 t 190 ppm NO at a flow rate of 2 4.9 l/min and seeded with 19% 0 t 4.76% CO2 t 0.23% NH3 t 76% 2 N2 flowing at a rate of 0.84 I/min [41]. The time to reach a complete removal of NO, depended on the applied voltage level. At 28 kV it took 1 h to reach complete removal (100%) of NO,while at higher voltage the steady state condition was reached at a much faster time (- 10 min) (Figure 24) [41]. The removal of CO2 from a simulated gas mixture of Nz: 0 : 2 CO2 = 0.744 0.158: 0.98 at l.01x105Pa was investigated using dc corona with the addition of argon [57]. The discharge reactor employed a pair of cylindrical hollow electrodes through which the gas mixture flowed [52,58]. The removal of CO2 increased with increasing corona discharge current and with the addition of argon (Figure 25) [57]. It also increased with addition of an argon admixture to 0.4 (ratio of argon to gas mixture) and thereafter decreased with a further increase in the argon content to a ratio of 2.4 (Figure 25). The loss of CO2 was accompanied by a corresponding increase in the concentration of CO which is partly produced by collision of excited atomic oxygen with CO2 [57,59].The production of CO in relation to CO2 depended on the discharge current and it increased with increasing current [57].

7 REMOVAL OF POLLUTANTS USING ARC PLASMAS


A few studies have been conducted on the removal of NO, and SO2 from simulated exhaust gases by creating an arc plasma using argon

664

Hackam et al.: Air Pollution Control by Electrical Discharges

Power k W

Figure 27. NO, Removal ratio as a function of dc power of plasma jet in N2 and for different 0 2 concentrations. 2150 ppm of NO in the simulated gas mixture; other conditions are as in Figure 26 [64].
Id(mA1

Figure 25. Reduced CO2 from simulated combustion gas as a function of dc corona discharge current for different ratios 7 of argon to simulated gas mixture. Gas mixture: N2: 02: CO2 = 74.4 15.8: 9.8; 7 = Ar/(N2+02tC02); total flow rate 0.7 I/min pressure l.01x105 Pa.
[60,61], nitrogen [62,63] and with N2 arc plasma injection into the simulated gas mixture [64,65]. A typical NO, removal ratio in a mixture 2 of 346 to 2317 ppm of NO, to 11% 02, to 8% COX, and the balance N with an undetermined amount of water vapor flowing at 20 l/min and

2 a dc arc plasma jet in nitrogen for different amounts of 0 content in the range 3.3 to 10.23% [64]. The initial concentration of NO was 2150 ppm while all other conditions remained as for Figure 26 [64]. It will be observed from Figure 27 that the maximum reduction of NO, occurred at a low power of arc plasma of 600 W and the removal ratio decreased with increasing power and increasing oxygen content. With larger power, NO, increased rather than decreased, and Figure 27 shows this as negative removal ratios (64).
N

a plasma jet generated in nitrogen and injected into the mixture containing the NO at a flow rate of 13 l/min is shown in Figure 26 as a function of dc plasma power [64]. The dc arc plasma (20 to 35 A, 15 to 60 V) was generated using thoriated tungsten electrodes. The plasma generates N radicals which remove the pollutants. Figure 26 shows that when the plasma power is in the range 0.8 to 1.2 kW the NO, removal ratio is at 100%. At higher powers the temperature of the plasma was reported to increase and the removal ratio of NO, decreased (Figure 26) and this was attributed to the regeneration of NO, [64]. At low powers ( 4 . 6 kW) the plasma arc was unstable, resulting in a lower removal of NO, (Figure 26) and this was attributed to incomplete dissociation

REMOVAL OF POLLUTANTS USING ac

Discharge reactors for pollution control using dielectric barriers (Figure 3@)) [23,28,67,69,70-72,74771, with dielectric pellets (Figure 3(c)) [66,68,73,75] or surface discharges [23,28] often employ alternating current (ac) at the power frequency of 50 and 60 Hz or at higher frequencies. Dielectric barrier [36] and surface discharge [39] reactors also have been used with pulsed power, but these will be discussed later in this paper (Section 9).

100

200

300

400

500

Inlet NO Coocenvawn (ppmv)

4"

"

1 I

'

"

" 2

'

"

Power k W

Figure 26. Dependence of NO, removal ratio on injected input power in nitrogen plasma to a simulated flue gas. Gas mixture: 346 to 2317 ppm NO, to 11%02; to 8% Cor, balance N2 and undetermined amount of H20; flow rate of gas mixture 20 l/min; flow rate of N2 plasma jet 13 l/min; plasma arc 20 to 35 A, 15 to 60 V [64]. Figure 27 shows the NO, removal ratio as a function of power of

Figure 28. Dependence of the removal ratio (solid line) and the absolute (dashed line) of NO on its initial concentration using ac (60 Hz) in a dielectric barrier discharge reactor. Gas mixture: N2: 02: H20 C Q = 76: 6 6 12; pressure l.01x105 Pa; temperature 130C; discharge reactor: concentric cylindrical 2.4 mm central rod, quartz 40 mm in diameter, 2 mm thick; flow rate 2.7 I/ min; applied voltage constant at 23 and 25 kV [67]. Using a cylindrical quartz tube of inner diameter of 40 mm, and 2 mm thick with 2.4 mm diameter central rod, a dielectric barrier dis-

IEEE Transactions on Dielectrics and Electrical Insulation


charge was formed with a stainless steel wire mesh screen on the outside [67]. 60 Hz was used to initiate a discharge in a simulated gas of N2: 02: H20: CO2 = 76: 6: 6 12 while the NO concentration was varied from 150 to 500 ppm. Figure 28 shows the removal ratio of NO (solid line) as well as the absolute values removed (dashed line) using ac (60 Hz) as a function of the initial concentration of NO [67]. The input power into the discharge was kept constant for two fixed applied voltages of 23 and 25 kV. A 94% removal ratio of NO was obtained at 25 kV applied voltage at the lowest initial concentration of 150 ppm of NO but this removal ratio decreased to about 58% at 500 ppm. Figure 28 also shows that both the absolute number of the removed NO and the removal ratio increased slightly with increasing ac voltage at a fixed power deposition into the discharge. It was also found that the NO removal ratio decreased slightly with increasing concentration of NO at a fixed applied voltage and increased with increasing H20 [67].
800

Vol. 7 No. 5, October 2000

665

- 1
s
E

500

2
U

400

NOx

300

IO0

I, ,;'. ';
.'
0

'.--_

NO

----------*
; N O 2

...........

..............................

10

700

power Iwi

600
"-7

Figure 30. As for Figure 29, except for out-of-phasecombined surface and dielectric barrier discharges [28].

e
U1

500

3 2 I
U

400

5
.0 c

200

150

300
200

%
NO2
. P
0

100

too
0
0

,.*"

_........*.........................

50

0
2 4

10

10

100

power &Vl

Input energy density [J.L-'I

Figure 29. Dependence of NO,NO2 and NO, on ac input power for in-phase and combined surface and dielectric barrier discharges. Discharge reactor is shown in Figure 4. Gas mixture: NI t CO2 t 0 (pro2 portions were not specified)800 ppm NO; flow rate 2 b i n ; pressure l.01x105 Pa [28].

A discharge reactor (Figure 4) containing a dielectric barrier and a surface discharge was used with ac (60 Hz) to study the removal of NO, with the two discharges combined either in-phase or out-of- phase [23, thick as a dielectric barrier and up to three blocks of needle electrodes, 281. Figures 29 and 30 show the dependence of the concentrations of each comprised of 275 needles of 13mm in height and 1.4 mm in diameNO, NO? and NO, on the input power for in-phase and out-of-phase ter set on a holder base of 84x 58 mm2. The flow rate was 5 to 20 l/min. applied ac voltages, respectively [28]. A simulated gas mixture con- Figure 31 shows that the initial concentration of NO of 200 ppm in a taining an initial concentration of 800 ppm of NO and NP t CO2 t 0 mixture of Nz:2 = 9: 1 decreased with increasing input energy into 2 0 (proportions were not specified) was used [28]. It will be observed the discharge and it was independent of the number of the needle electhat there are large reductions in the concentrations of NO and NO, trodes [70]. Figure 31 shows that it was possible to remove NO comfor the out-of-phase (Figure 30) compared to the in-phase (Figure 29) pletely when the ac input energy was >150 J/I [70]. discharges. The reduction of NO, from the exhaust gases of a diesel engine was Figure 31 shows the dependence of NO and NO, removal on the investigated using 50 Hz power applied to a cylindrical barrier disinput energy (J/l) to an ac dielectric barrier discharge operating at 10 charge reactor [71]. The reactor comprised a stainless steel rod 6 mm in to 100 kHz [70]. The discharge reactor employed a flat soda glass 2 mm diameter and was 300 mm long. The outer electrode was an aluminum

Figure 31. Dependence of NO and NO, removals on input energy density to an ac (10 to 100 kHz) dielectric barrier discharge. Initial concentration of NO 200 ppm; gas mixture 90% Nz, 10%02; pressure l.01x105 Pa; temperature 10 to 25C; dielectric barrier, flat soda glass 2 nun thick; each block A-type electrode (x 1) = 275 needles, 13 mm high, 1.4 mm in diameter on a holder base 84x 58 mm2;up to 3 blocks used; flow rate 5 to 20 l/min 1701.

666

Hackam et al.: Air Pollution Control by Electrical Discharges

0 PukedCorons
0

ACDielecMcRanier

P
5 6 7 0 9 1011 1 2 0 1 4 1 5 Applied voltage (kV)
0324 153 Input power [kVAl 0
2.88

4g
0.1
0.01

100 200 300 400 Energy Density (JA)

600

Figure 32. NOx reduction as a function of applied ac (50 Hz) voltage and input power into a coaxial Pyrex glass dielectric barrier discharge reactor. Exhaust gases from a diesel engine. Curve 1,5 discharge reactors per transformer; curve 2,lO discharge reactors per transformer
[711.

Figure 33. Dependence of the removal ratio of the volatile organic compound (VOC) trichloroethylene (TCE) in dry air 8s. energy density Discharge reactors: flat bed dielectric barrier (1.2 kHz) and pulsed corona discharge reactors; initial concentration of TCE 200 ppm; dielectric reactor: flat glass 3 mm thick, 38 cmx 70 cm, gap separation 3.5 mm; pulse corona reactor: coaxial, stainless steel electrodes, 0.5 mm wire diameter, 25 mm cylinder inner diameter, 90 cm long; pulse voltage: rise time 6 ns, FWHM 20 ns, peak voltage 26 kV; X final concentration of vOC, X , initial concentration [72].

foil placed on the outer surface of a Pyrex glass tube, 1.6 mm thick and 16.8 mm in inner diameter [71]. Figure 32 shows the dependence of the removal ratio of NO, on the applied ac (50 Hz) voltage and input power into the discharge in exhaust gases from a diesel engine [71]. The removal ratio of NO, increased with increased voltage and increased input power. Furthermore, a higher removal ratio was obtained when the input impedance of the transformer was smaller at 31.25 kR (Figure 32, curve 1)compared to 62.5 kR (Figure 32, curve 2) [71]. The effects of SO2 concentration on SO2 removal ratio was studied using 60 Hz at 24 kV peak to peak in a coaxial dielectric barrier disz 2 charge reactor employing glass [69]. A mixture of S02, N ,and 0 at a pressure of 8 . 7 ~ 1 Pa~at room temperature was used. The SO2 con0 centration was in the range 300 to 14000 ppm. The highest reduction ratio was 70% which occurred at low flow rate and this decreased to 15%at a high flow rate. Increasing the pressure at the same applied voltage resulted in a reduction in the removal of SO2 due to the lowering of the electric field to the pressure ratio ( E / P ) causing a lowering of the electron density. Similarly, lowering the applied voltage for the same gas pressure and fixed flow rate resulted in a lowering of the SO2 removal [69]. A dielectric barrier discharge reactor employing two flat plates glass (3 mm t h i c k 38 cmx 70 cm) with a gap spacing of 3.5 mm and aluminum electrodes with 1.2! d z ac driven power source (3 kW) was used to remove three volatile organic compound (VOC) in dry air [72]. The voc removal of trichloroethylene (TCE), methylethylketone (MEK) and methylene chloride (MECL) were studied over a wide range of energy density input to the discharge using both ac and pulsed voltages and are shown in Figures 33 to 35 [72]. The pulse corona reactor employed a concentric coaxial stainless steel electrodes with a wire of 0.5 mm in diameter, 25 mm of inner diameter of the outer cylinder and 90 cm long. A positive pulsed voltage having a rise time of 6 ns, a duration of full-width-at-half-maximum (FWHM) of about 20 ns and a peak voltage of 26 kV. The results from the pulsed corona are also presented

0.1

ACDielecMcBanier

0 . 0 1 ~ " " " " ' " " " ~ ' " ~ 0 1 2 3 Energy Density (mi)

Figure 34. As for Figure 33 except for the removal of methylethylketone (MEK).Initial concentrationof MEK in air 1000 ppm [72]. in figures 33 to 35 together with the removal ratios using ac [72]. The removal ratios of all three voc shown in Figures 33 to 35 decrease with increasing energy density u/1) input into the discharge for both the ac and pulsed driven power sources [72]. Figures 33 and 34 show that the removal ratio X / X o ( X = final concentration, X, = initial concentration of VOC) for TCE and MEK, respectively in dry air mixture were similar for both types of energy density of either ac or pulsed. Figure 35 demonstrates that for MECL in air also similar energies were required from either sources up to 1kJ/l but for higher energy density the pulsed corona discharge had significantlybetter removal efficiency. The difference in the removal efficiencies shown in Figure 35 was attributed to the increased temperature of the gas. This was measured and is shown for the pulsed corona reactor in Figure 35. The removal efficiency was also independent of the concentration for 100 to 200 ppm of TCE and for 200 to 1000 ppm Of MECL [72].
TCE in air for ac (1.2 kHz) and pulsed

Figure 36 shows a comparison between the removal efficiency of dielectric barrier discharges [72].

XEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7 No. 5, October 2000

667

100
0
ACDbkctricRanler

-e

: 80 2 A 3 60
+
0

0.01 0

' ' ' ' ' ' ' I' ' ' ' I ' ' ' ' "' ' ' I' ' ' ' I ' ' ' ' A

'

5 .- 40 2
a ,

>

2 3 4 8 6 Enem Density (kJ/I)

20

Figure 35. As for Figure 33 except for the removal of methylene chloride (MECL). Initial concentration of MECL in air, 1000 ppm; temperature in "c is shown for the pulsed corona discharge reactor [72].
0
0

n - 0

4 6 8 Input power I W
( 1 ~
~ ~ ~~

10

Puked DbbcMcBnler ACDbkctrlcBnler

Figure 37. Dependence of the removal of benzene in dry air on 50 Hz inout Dower in a ferroelectric I oellet discharee reartor usiw differ... ..0 -~~~~~~ .. ~~-~~~~ ... en't peimittivities of the pellets. Gas mixture: 80% Nz, 20% 0 ; 2 dis~~~~ ~ ~ ~~~~

charge reactor: cylindrical, rod diameter 10 mm, inner diameter of outer cylinder 30 nun, 50 mm long; pellets diameters 1 to 3 mm; pellets held between perforated Teflon plates; gas flow rate 0.2 I/min; initial concentration of benzene 200 ppm. Relative permittivities: curve 1,20 (MglTiOa);curve 2,100 (CaTiOa),curve 3,200 (CaTi03);curve 4, 330 (SrTiO3);curve 5,870 (BaTi03);curve 6,1100 to 15000 (SrTiOs and BaTi03) [66].
0
100 200

300

400

600

Energy Density (J/I)


Figure 36. As for Figure 33 except for pulsed and ac driven dielectric barrier discharges. ac coaxial dielectric barrier reactor as in Figure 33; pulsed dielectric barrier reactor: a singlebarrier 3 mm thick flat Pyrex glass, 38 mm in diameter, gap spacing2 mm; stainless steel electrodes: negative peak voltage 40 kV; rise time 6 ns; FWHM 20 m;initial concentration of TCE 200 ppm [72]. In this case the 3 mm thick Pyrex glass dielectric barrier covered one flat electrode 38 mm in diameter and forming a gap spacing of 2 mm. Figure 36 shows that the removal efficiency of TCE in dry air was the same for pulsed dielectric barrier and ac dielectric barrier discharges for up to 320 J/I [72]. The higher breakdown field in the pulsed dielectric barrier discharge compared to the ac dielectric barrier discharge did not affect the removal efficiency for TCE in air (Figure 36) [72]. Benzene (C6H6) in air was decomposed in a ferroelectric pellets coaxial reactor using either 50 Hz or 24 kHz power sources [66]. A cylindrical reactor was used with a central electrode 10 mm in diameter, outer electrode of 30 mm inner diameter and 50 mm long. The pellets were 1 to 3 mm in diameter and were held in position between perforated Teflon plates [66]. It was found that ferroelectric pellets of 1 to 2 mm in diameter and having a relative permittivity >1100 (BaTiOs and SrTiOs) decomposed benzene at an input power of 8 W (Figure 37) [66]. The benzene in dry air (80% N2 and 20% 02)was completely decomposed into CO and CO2 at low concentrations (t50 ppm) without formation of other hydrocarbons [66]. It also was reported that the use

of 50 Hz resulted in a higher energy efficiency for the decomposition of benzene compared to 24 kHz, and further that the concentrations of NO, and nitrous oxide (N20) were lower with the lower frequency power source [66].

-0- 5 7 p m m + mipppm.. l13mmm pm mm + n9pP m s M n +6

+ P.ppm.5"
0

10

12

14

18

ac Voltage (kv)

Figure 38. Destruction efficiency of toluene us. ac voltage (60 Hz) using a ferroelectricpellet discharge reactor. Initial concentrationsof toluene and pellet diameters are shown. BaTi03 pellets; gap separation between the flat mesh electrodes 25 nun; flow rate 0.2 l/min; gases: dry air, oxygen and nitrogen [73]. Figures 38 and 39 show the destruction efficiency of toluene (C~HB) , and trichlorotrifluoroethane ( c ~ ) - 1 1 3respectively, using 60 Hz applied voltage for a flow in the range of 0.1 to 0.5 l/min (dry air, oxygen and nitrogen) and for varying initial concentrations from 57 to 234 ppm and using BaTiOa pellet diameters in the range 1to 5 mm [73]. The gap separation between the flat mesh electrodes which held the pellets in position was 25 m.Figure 38 shows that the destruction rate depends

668

Hackam et al.: Air Pollution Control by Electrical Discharges


1. lower cost of the equipment, 2. absence of high energy X-rays and other radiation which might affect the personnel working in the treatment area, 3. homogeneous discharges at atmospheric pressure can be produced, 4. high energy electrons are generated readily, which produce the radicals that are necessary for pollutant removal, and 5. because the pulsed power is very short in duration, the ions and neutrals are not heated significantly above the ambient [lo, 111.
15

I-

-m-

500 ppm, 0.1 Umin -+ 500 ppm, 0 2 Umin -A500 ppm, 0.5 Umjn o 1000 ppm, 0.2 Umm

The homogeneity of the pulsed discharge was confirmed recently using 25 ns pulse which generated negative corona in air containing Voltage (kv) NO in a needle to plane gap geometry with a flowing gas [92]. A uniform destruction of NO (40% removal ratio) throughout the gap was Figure 39. Destruction efficiency of trichlorotrifluoroethane (CFC)-113 found, thus indirectly confirming that the discharge was homogeneous as a function of ac voltage (60 Hz) in a pellet discharge reactor. Initial concentration and flow rates are shown in the Figure. Pellet size 3 mm; throughout the electrode gap. other conditions are as in Figure 38 [73]. Concentric wire-cylinder discharge reactors of the type shown in Figure 3(a) [36] and Figure 8 [32] have been used widely in pulsed on the pellet size. This is because the corona onset voltage decreased as the pellet diameter increased [73]. A 100%destruction rate of toluene power operations with some notable improvement by modifying the was attained using 60 Hz applied voltage for all pellet diameters in the HV electrode from a straight wire to helical construction as shown in range 1 to 5 mm (Figure 38) and higher efficiencies were obtained at Figure 9 [42]. Using the three types of cylindrical discharge reactors shown in Figlower initial concentrations of toluene (57 and 59 ppm) than at higher ure 3 and employing short duration pulses of 100 ns in a mixture of concentrations (231 to 237 ppm) (Figure 38) [73]. Figure 39 shows the destruction efficiency of CFC-113 as a function 100 ppm NO in nitrogen with a gas flow rate of 65 l/min, it was found of the applied ac voltage using 3 mm diameter BaTi03 pellets and for that the energy density input to the discharge necessary to remove NO different initial concentrations in the range 500 to 1000 ppm and flow generally was similar in magnitude in all three reactors for 5 50 J/1 rate from 0.1 to 0.5 l/min [73]. The maximum destruction efficiency was 1361. It was concluded that the production efficiency of the radicals per 28% with a flow rate of 0.2 l/min and 500 ppm. The low destruction efficiency of CFC-113 was due to the stronger bonding of the chlorine input energy in electrical discharge reactors could not be increased by and fluorine in the molecule compared to toluene [73]. When smaller changing the voltage pulse parameters, changing the electrode strucpellets of 1 mm in diameter were used, a higher operating voltage of tures, employing a dielectric barrier, or dielectric pellets [36]. This was attributed to the field in the plasma producing the radicals, being space 24 kV,, could be realized with a destruction efficiency of 52% [73]. Formaldehyde and benzene also were removed from gas streams of charge shielded to the same value in all cases [36]. oxygen using a 60 Hz dielectricbarrier discharge reactor [75]. The initial 10 EFFECTS OF VOLTAGE concentrations of the pollutants were in the range 100 to 5000 ppm and POLARITY OF PULSED the flow rate from 0.1 to 1.0 ]/min. The discharge gap was 2.5 mm POWER ON PERFORMANCE and the length of the cylindrical dielectric barrier discharge reactor was varied between 10 to 100 cm [75]. It has been shown that in a mixture of 180 ppm NO in air at room temperature the application of positive polarity pulses to the wire of a 9 POLLUTION CONTROL USING coaxial discharge reactor (Figure 3(a))produced a better performance in PULSED POWER the removal of NO than negative pulses, although the energy efficienPulsed power is increasingly being used for removal of pollutants cies in both cases were the same [4]. However, at higher temperatures the situation was reversed and the negative polarity yielded a better such as NO, and SO2 from flue gases [3,4,7,10,11,21,29,31,39,50,78871 as well as volatile compounds such as toluene (C7H8) [73,87,88], performance. In a combustion gas from an incinerator boiler plant both dichloromethane (CHzC12) [88,89], dichlorodifloromethane (CFzC12) polarities gave a similar performance while the energy efficiency was [88], trichloroethylene [72], methyl ethyl ketone [72], methylene chlo- better for the positive pulse polarity [4]. The difference in the perforride [72,108], sulfurhexafluoride [89,94], perfluorinated compounds mance due to changing the voltage polarity in the flue and the sim(NH3, C2F6 and CFa) [89], chlorofluorocarbons (CC12F2, C7H8 and ulated gases used in the laboratory was attributed to the presence of CH3CC13) [89,94], trichlorotrifluoroethane (CFC-113) [39,73], methane 17% H20,10% COz, 300 ppm CO and 350 ppm HC1 as well as due to [90], hydrogen sulfide (HzS)[91],methylene chloride (CH2C12) [73,94], the presence of fine particulates in the combustion gases. styrene [87], ethylene [87], CF4 [94], C2F6 [94], NF3 [94] and carbon Figure 40 shows a comparison in the reduction of NO and NO, for tetrachloride (CCla) [108]. positive and negative polarities as a function of peak pulsed voltage The main advantages of using the pulsed power method for the re- using FWHM = 350 ns in a simulated gas mixture of 6% 02, 8% CO2, moval of pollutants from flue and exhaust gases compared to the elec- 7.5% HzO, 78.3% Nz, 0.2% SOz, 205 ppm NO, 35 ppm NO2 and 0.8 tron beam irradiation method are: stoichiometric NH3 (NH3/(NOXt SO2)) [86]. The discharge reactor
*O 0 0
5
1 0
20

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7"o. 5, October 2000

669

Generally the shorter the rise time of the applied pulse voltage, the more intense and elongated are the pulsed streamers and the better the performance [4].

12 EFFECTS OF PULSE WIDTH AND REPETITION RATE ON REMOVAL OF POLLUTANTS


It was reported that the use of dc voltage produced a large ionic current which reduced the efficiency of the discharge reactor and therefore it was concluded that a very short pulse width would be advantageous [4]. A reduction in the pulse width from 165 to 81 ns and to 41 ns with corresponding shorter rise times of 25, 15 and 9 ns resulted in an increase in the destruction of NO (100 ppm) in nitrogen and toluene (C7H8) in air [94]. Figure 41 shows the percentage removal of toluene in air for different pulse lengths as a function of the specific energy input into a coaxial pulsed corona discharge reactor [94]. It will be observed from Figure 41 that a higher destruction rate was obtained at a shorter pulse width for a given energy density input, and therefore a higher energy efficiency. Figure 41 also shows that in order to attain a required destruction ratio less energy would be required when decreasing the pulse width from 165 to 45 ns 1941.

Fieure 40. Deuendence of the removal of NO and NO, on the uolaritv " I , and magnitude of pulsed voltage in a coaxial wire-cylinder geometry Gas mixture: 6% 02,8% C02,7.5% H20,78.3% Nz, 0.2% S02,205 ppm NO, 35 ppm NO>, 0.8 stoichiometric NH3; gas flow rate 14 Vmin; room temperature; FWHM 350 ns; discharge reactor: coaxial, 3 mm wire diameter, 55 mm inner diameter of outer cylinder, length 45 cm; pressure l.01x105Pa 1861.

used was a coaxial reactor, 3 mm wire diameter, 55 mm inner diameter of the outer cylinder and 45 cm long. It will be observed from Figure 40 that the positive polarity pulses were more effective in removing NO andNO, [86]. For a mixture of air with 2.5% H20 and 1666 ppm SO2 and a flow rate of 1.2 I/min in needle to plate or rod-plate geometries, the positive pulsed voltage performed better than the negative polarity in removing SO2 [7]. This was because the positive pulse polarity produced more uniform streamers that extended further into the gap and 90% of SO2 was thus removed with positive polarity [7].

75

!-

Pulse Width FWHM

165 ns

300

400

500

600

700

EO0

Specific Energy [JA]

11 EFFECTS OF MAGNITUDE OF PEAK PULSED VOLTAGE ON PERFORMANCE

Figure 41. Destruction of toluene in air using different positive pulse widths in a coaxial corona discharge reactor. Pulse rise time and corresponding FWHM: 9 ns and 45 ns; 15 ns (81 ns); 25 ns (165 ns); concentration of toluene 200 ppm [94].

It has been found that the removal ratio of the pollutants increases Figure 42 shows the removal efficiency of NO (mol/kWh) as a funcwith increasing peak value of the pulsed voltage when all other condi- tion of the removal ratio of NO in a mixture of 91% N2, 5% 02, 4% tions are kept unchanged [4,81]. A typical dependence of the removal HzO and an initial concentration of 200 ppm NO for different pulse of NO and NO, in a mixture of 6% 02, 8% COz, 7.5% H20,78.3% Nz, widths from 40 to 120 ns [43]. It will be observed that a higher removal 0.2% SO2 and 240 ppm NO, on the peaks of positive and negative volt- efficiencywas obtained with successively shorter pulse widths. In Figages is shown in Figure 40 [86]. It can be seen that both NO and NO, ure 42 the pulse rate increased from 1to 13pulses per second (pps) with concentrations decreased with increasing peak voltages for both po- successive points on each curve starting from lower to higher removal larities but the rate of decrease was more pronounced with increasing ratios of NO [43]. positive pulsed voltage (Figure 40) [86]. The removal of NH3 and 0.5 Figure 43 shows that at a fixed pulse width the final concentration to 1.0 p m particulates in air, also was found to be enhanced with in- of NO decreased with increasing pulse rate. For a pulse width of 120 ns creasing the peak pulse voltage [93]. Masuda reported that the removal the final concentration of NO decreased from the initial concentration of NO in a mixture of 180 ppm NO, 600 ppm NH3 in air using nega- of 200 ppm to 4 ppm at 10 pps which was 98% removal efficiency [43]. tive pulsed voltage was enhanced with increasing the magnitude of the At a fixed pulse repetition rate, the final concentration decreased with peak of the pulse [4,81]. increasing pulse length. Typically at 7 pps the final concentration of

670

Hackam et al.: Air Pollution Control by Electrical Discharges

9
L-

NO removal ratio, NOR, %

20

40

60

80

100

Figure 42. Removal efficiency of nitric oxide (mol/kWh) as a function of its removal ratio using different pulse widths. Gas mixture: 91% N2, 5% 02, 4% H 0 and 200 ppm NO; flow rate 2 l/min; pressure 2, l.01x105 Pa; temperature 25T; discharge reactor of the type shown in Figure 8; 0.5 mm diameter stainless steel wire, 76 mm inner diameter cylinder, 0.5 m long; positive peak voltage 49.2 kV; pulse rate increases from 1 to 13 pps with successive points on each curve from 60 lower to higher removal ratios; 0 40 ns; I ns; 80 ns; A 100 ns; v 120 ns [43].

Figure 44. Dependence of the removal ratio of NO and NO, on the pulse repetition rate in a mixture of 10% 02, 2% H20, 88% Nz and 200 ppm. Positive voltage pulse, FWHM 120 ns; flow rate 1.94 l/min; coaxial discharge reactor: 1 mm wire diameter wound spirally at 1 mm pitch on 25 mm diameter insulating tube, 96 mm inner diameter of outer cylinder, 3.5 m long; pressure l.01x105Pa 1421.

13 TIME TO STEADY STATE CONDITION WITH PULSED POWER


The time to reach a steady state condition of the final concentration of the pollutant depends on many parameters including the flow rate of the gas, the concentration of the pollutants, the composition of the gas mixture, the input pulse power to the discharge and the size of the discharge reactor. There is a dearth of published results on the time to reach a steady state condition. Figure 45 shows the final concentration of NO and NO2 as a function of the time of application of a positive pulsed voltage of 49.2 kV, 120 ns pulse length and 1pps to a gas mix2 ture containing 200 ppm NO, 91% Nz, 5% 0 and 4% H20 at a flow rate of 2 l/min [43]. It will be observed that it took 3 min for the concentration of NO to decrease from 200 ppm to 140 ppm and then remained constant thereafter (Figure 45) [43].

4 (a)

E 1 0 1 2 1 4

Pulse repetltlon rate, pps

E 1
c 100

Pulse repetltlon rate, pps

8 1 0 1 2 1 4

Figure 43. Reduction of NO (a) and NO1 (b) as a function of pulse repetition rate using different pulse widths. Conditions and symbols are as in Figure 42 [43].

a l
' 0

Figure 45. Dependence of NO and NO2 concentrations on time of to 120 ns (Figure 43) [43]. application of positive pulsed voltage. Pulse rate 1pps; FWHM 120 ns; In a mixture of 10% 02,2% H20,88% N2 and 200 ppm NO and a other conditions as in Figure 42 [43]. flow rate of 1.94 l/min, NO was removed completely at 3 pps while 96% Figure 46 shows the concentration of NO, as a function of time folof NO,was removed at 7 pps using a coaxial reactor having a spirally wound 1mm wire in diameter, 96 mm inner diameter of outer cylinder lowing the application of positive pulses of FWHM = 20 ns to a simulated gas mixture of 400 ppm NO,10% 02,10% CO2 and 80% N2 at a flow and 3.5 m long (Figure 44) [42].

NO decreased from 49 to 24 ppm with increasing pulse length from 40

Time, min

IEEE Transactions on Dielectrlcs and Electrical Insulation

Vol. 7 No. 5, October 2000


1M) I

671

400
Y

--c Spray rcactor -t- Wet rcocior - 1 4 pH value

'2 3M)

E 8
,^

80 60 -

with discharge @eNO) with discharge (DeNOx)

-.

6
0

z :
20

40-

*\

10 20 Voltage application timo [min]

30

z o

i .

Figure 46. Dependence of NO, concentrationon the time of application of positive pulsed voltage. Gas mixture: 400 ppm NO, 10% 0 2 ; 10% CO2; 80% N2; flow rate 2 l/min; pressure l.01x105 Pa; temperature 25C; input power 30 W; discharge reactors: coaxial glass dielectric barrier 1mm thick, 23 mm inside diameter, 210 mm long; wet reactor, 0.2 mm diameter stainlesssteel wire; spray reactor: 2.5 mm diameter tube with holes (0.15 mm diameter) forming the HV electrode [951. rate of 2 l/min using a dielectric barrier discharge reactor [95]. It will be observed that a saturation is reached after 4 to 15 min of application of the pulsed voltage, depending on the experimental conditions. In Figure 46 the pH value of HzO due to the absorption of NO2 is also shown where it took 30 min to reach a steady state value [95].

Figure 47. Dependence of the removal efficiencies of NO and NO, on the gas flow rate using positive square pulsed voltage. Gas mixture: 940 ppm NO, 10%02, 89.9%N2; discharge reactor: coaxial dielectric barrier 0.2 mm diameter wire, quartz 1.5 mm thick, 20 mm inner diameter; room temperature; 60 pps [96].

56 kV(11.8

kV/cm)

14 DEPENDENCE ON FLOW

RATE AND RESIDENCE TIME


Figure 47 shows the NO and NO, removal efficiency as a function of the flow rate of a simulated mixture of gases (940 ppm NO, 10% 02, and 89.9% N )using positive square pulses which were applied to a z dielectric barrier discharge reactor [96]. The coaxial discharge reactor was made of 0.2 mm diameter wire concentric in a 20 mm quartz tube 1.5 mm thick, which was wrapped on the outside with an aluminum foil [96]. It will be observed from Figure 47 that at room temperature respectively the removal efficiencies of NO decreased from 30% to lo%, with increasing flow rate from 1 to 8 I/ min [96]. Figure 48 shows the change in NO concentration as a function of residence time in a coaxial discharge reactor for different negative pulsed voltages and pulsed electric fields [U]. It will be observed that the removal ratio of NO in a mixture of 180 ppm NO, 600 ppm NH3 in air decreased with increasing residence time in the discharge reactor for all applied voltages (Figure 48) [81].

c I

" >
30

2 0

60

G a s Rertdence T t n e Ta ( 5 )

Figure 48. Change in NO concentrationas a functionof residence time in a coaxial discharge reactor for different negative pulsed voltages and pulsed electric fields. Discharge reactor: coaxial, wire 3 mm diameter, cylinder 100 mm inner diameter, 416 mm long; pulsed voltage: rise time 50 ns, FWHM 300 ns, pps 50; gas mixture: air with 180 ppm NO 600 ppm NH3; pressure 1 . 0 1 ~ 1 Pa; temperature 20C [81]. 0~ 72,95-971. Generally, higher gas temperatures lead to a lower removal efficiency of the pollutants. The application of pulsed power for the removal of pollutants does not raise the temperature of the gas significantly above the ambient. This was confirmed using spectroscopic measurements which showed that for streamer pulses lasting 400 ns, the gas temperature in atmospheric nitrogen remained in the range 300 to 350 K [loo, 1011. Figure 49 shows the magnitude reduction in NO concentration as a function of temperature in the range 20 to 220C in a mixture of 400 ppm NO, 10% 0 and 90% N using positive pulsed voltage in a coaxial di2 z electric barrier discharge reactor [9]. Different input power levels into the corona discharge were employed. It will be observed that an increase in the gas temperature resulted in a decrease in the NO removal (Figure 49) [9]. For example the decrease in NO concentration at 20C was 200 ppm while at 220C only 76 ppm of NO were removed [9].

15 EFFECT OF GAS TEMPERATURE ON THE REMOVAL OF POLLUTANTS


The temperature of the flue gases is usually reduced by heat exchangers to 80 to 100C [97-991 before they are treated either in a corona discharge or electron beam reactors. The exhaust gas temperature from a diesel engine was reported to be in the range 150 to 200C [96]. The effect of varying the temperature of the gas on the removal of the pollutants has been reported in several studies [4,9,14,67,

- -

672

Hackam et al.: Air Pollution Control by Electrical Discharges

150

ppm

- 80 s

400 600

ppm

ppm

0 0

01

1
100

200

Tempemture ["Cl Figure 49. Reduction in NO concentration as a function of gas temperature for different input power into the discharge. Reactor: coaxial dielectric barrier, glass 1mm thick, 18mm inner diameter, 78 mm long; positive pulsed voltage, width 500 ns; gas mixture: 400 ppm NO, 10% 02,90% N2; flow rate 2 l/min; pressure 1.01x105Pa [9]. Figure 35 shows that for a dielectric barrier discharge using pulsed power the removal of methylene chloride (MECL) in dry air deteriorated with increasing temperature from room temperature to 60C [72]. The percentage removal of NO in a mixture of 940 ppm NO, 10% 0 2 and 89.9%N2 using positive pulsed voltage in a dielectric barrier discharge reactor increased when the temperature decreased from room temperature to -114C (solid ethanol) and to -196C (liquid nitrogen) [96]. This occurred over a wide range of gas flow rate from 1 to 8 ]/min. Typically at 4 I/min flow rate, the removal ratios of NO were 15.7%,60.6%,and 98%, respectively at room temperature, -114C and -196C [96]. The removal of NO in air using negative pulsed voltage was less effective with increasing gas temperature from 50 to 250C [4]. For a positive polarity pulsed voltage at 250C not only NO was not removed, but its concentration actually increased above the initial value and likewise at 200C for peak voltages >60 kV [4]. The removal ratios of SO2 in a mixture of 73% N2, 13% C02, 6% 0 and 8% H20,350 to 550 ppm NO,, 300 to 550 ppm SOX,a ratio of 2 NHs/(NO, t SO*)0.7 to 0.8 and using pulsed power (FWHM = 1.5 /AS, = pps = 50) and either parallel plates or coaxial discharge reactors, were reported to be 75% at 100C and increased to 90% with decreasing gas temperature to 70C [97].

100

.e

CFC-113
60
Frequency: 5kHr Pc'rkVollogc:6kV Eleclrlc Poycr
0

0"
0

16 EFFECT OF INITIAL CONCENTRATION ON REMOVAL EFFICIENCY


The concentration of the pollutants in the flue gases varies over a wide range depending on the type of fuel used. The concentration of NO, from coal burning plants varied from 300 to 550 ppm [97], 430 to 550 ppm [99], and 250 ppm [l,981 while those of SO2 varied from 350 to 550 ppm [97], 500 ppm [98] and 360 to 500 ppm [99]. The flue gas from a methane burner had 40 ppm NO and 0 NO2 [97]. In general at a given input power to the discharge, the removal ratio of the pollutants decreases with increasing initial concentration of the pollutant.

Cwcntntion: l00ppm * K W m

* WOPPm

2 4 6 8101214 Residence Time(s)

Figure 51. Decomposition of CFC-113 (freon 13) in air for different initial concentration using surface discharge reactor [39]. The destruction of CFC-113 (Freon-13)in air using a pulsed power for

IEEE Transactions on Dielectrics and Electrical Insulation


different initial concentrations in the range 100 to 104 ppm is depicted in Figure 51 [39]. It will be observed that for a given residence time in the surface discharge reactor a higher decompositionratio was obtained with a lower initial concentration of CFC-113 [39]. In flue gases from a coal burning power plant the NO, removal efficiency was also found to depend on its initial concentration [97].

Vol. 7 N o . 5, October 2000

673

17 INFLUENCE OF ADDITIVES ON THE REMOVAL OF POLLUTANTS


Additives are introduced into the gas stream of discharge reactors in order to enhance the removal of the pollutants and neutralize the nitric and sulfuric acids [2] which are produced by interaction of the radicals with NO, and SO2 (reactions(12), (14), (16), (18), (23) and (26)). One of the most widely used additives is ammonia [5,6,23,25,26,28, 95,97-99,1021 which converts the acids into ammonium sulfate and ammonium nitrate which can be utilized as agricultural fertilizers [21]. Other additives such as lime 1971, methane 126,271, ethylene (C2H4) [6, 9,85,90], propylene (CJH~) argon with ammonia mixture [25,26, [6], 491, CO2 with ammonia mixture [40], hydrocarbons such as Z-propane1-01 or 2-propanol [103], copper coated zeolite catalyst [85,103], H20 [81,95,96,102] and hydrated lime have been used in discharge reactors. The addition of ammonia to the flue gases from a coal-burning power station resulted in increases in the removal ratios of both NO, and SO2 [97]. The flue gas contained 420 ppm NO,, 360 ppm S02,73% N2,13% C02,6% 0 2 and 8% H2O. The flow rate was 500 m3/h and the gas temperature was 90C. The removal ratio of NO, increased from 15 to 33% and of SO2 from 14 to 7570, respectively, before and after the addition of 0.7 to 0.8 stoichiometric ammonia (NH3/(NOXt SOz)) when positive pulses (FWHMat 1 ps) were applied at an energy density of 6 Whim3 [97]. Without the application of pulsed voltage, the mere addition of ammonia resulted in the removal of 55% of SO2 [97]. The injection of hydrated lime using compressed air at a stoichiometric unity value resulted in an increase in the removal ratio of SO2 from 19 to 31%, respectively before and after the application of positive pulsed voltage [97]. The injection of hydrated lime had no effect on the removal of NO, [97]. Figure 52 shows the removal efficiency of SO2 in air containing 6% H20 at 65C and as a function of the ratio of the added NH3/S02 where SO2 = 1000 ppm [102]. A dc applied voltage at a current density of 10 mA/cm2 and energy density of 20 J/g were injected into the discharge using multi-pin cathodes in a point-plane geometry The length of the discharge reactor was 1 m, the gas flow rate 70 to 200 m3/h and the gas flow speed was in the range 70 to 200 m/s [102]. Figure 52 shows that the removal efficiency increases with increasing NH3 until saturation is reached. It was reported that in flue gas from a coal-burningplant containing 430 to 550 ppm NO,, 360 to 500 ppm SOz, a flow rate of 600 m3/h at 70 to 100C and using a positive pulsed corona coaxial discharge reactor, removal efficiencies of 60%for NO, and 80% for SO2 were obtained with energy input of 15 Wh/m3 with injection of ammonia of 1000 ppm [99].
N
N

Figure 52. SO2 removal ratio from air as a function of addition of the ratio of NH3/S02. Conditions: initial SO2 0.1%; 6% H20; negative dc applied voltage; current 10 mA/cm2; energy density 20 J/g; temperature 65T; discharge reactor: multipin needle to plane [102].

I
0

8
r"
z

0% Additional

0.1

0.15

0.2

NH,

gas flow (I/nin)

Figure 53. Removal ratios of NO, NO, and surplus NH3 as a function of the added NH3. Treated gas mixture: 80% N2,20% 02, and 190 ppm NO with flow rate 4.9 1/ min; additional gases: 80% N2, 20% 0 2 with flow rate from 0.05 to 0.2 l/min; positive dc voltage, electrode gap

5 cm [ 0 . 4]

The added ammonia into the flue gases should be consumed within the discharge completely, because any amount left unused will contribute to air pollution and thus must be removed. Figure 53 shows the amount of NH3 left in the gas stream after treatment in a discharge reactor of the type shown in Figure 7 as well as the corresponding reduction ratios of NO and NO, as a function of added NH3 [40]. The treated gas mixture contained 80%N2,20% 0 2 and 190 ppm NO with a flow rate of 4.9 I/min. Additional gases of 80% N2, and 20% 0 with a 2 flow rate of 0.8 I/min were fed into the treated gas via a pipe electrode (Figure 7). Additionally, Ar with NH3 (Ar: NH3 = 97 3) were fed into the treated gas at a rate of 0.05 l/min (Figure 53) [40]. A positive dc voltage was applied to the pipe electrode. It will be observed in Figure 53 that a complete removal of NO and NOxcouldbe

674

Hackam et al.: Air Pollution Control by Electrical Discharges


Figure 19 shows the concentrations of NO, NO1 and NO, with the addition of a fixed amount of NH3 of 219 ppm (NH3/(NO,), = 0.5) and Figure 20 with addition of a fixed amount of methane of 438 ppm (CH4/(NOX),= 1)using a negative dc voltage in a plane-plane geometry discharge reactor [26].
600
100
0
I .

480 400

without HzO. NHi

1,

1%
Y)

260

i zoo
160

4
H20 IO ~ 0 1 %NHI400 pp .

roo
50
0
0
10

20

30

40

dc or chaqlng vonpge bv

(b)

IO

Input power [W]

20

30

40

Figure 55 shows the effect of varying separately the amounts of the added NH3 and CHa on the reduction of NO, when either pulsed or negative dc voltages were applied [26]. Figure 55 shows that the addition of ammonia was more effective than methane for reducing NO, Figure 54 shows the NO and NO, removal ratios as a function of infor both NHs/(NO,), = CH4/(NOX), = 1 and 0.5. Further increasing put power using positive pulsed voltage in a mixture of 400 ppffl NO, the stoichiometry (ratio of additives to the initial value of NO,) led to 10%02,10% CO2 and 80% N2 for four conditions of with and without a larger reduction in the concentration of NO, (Figure 55) [26]. addition of NH3 and H2O [95]. The discharge reactor was a coaxial Figure 55 shows that the negative dc was more effective at the same glass dielectric barrier, 1mm thick, 23 mm inner diameter and 210 nun long with 0.2 mm stainless steel wire. The flow rate was 2 l/min and voltage than the pulsed voltage [26]. However, Figure 3 of [26] indicates the temperature of the gas 150C. It will be seen on Figure 54 that in that the pulse used was very wide such that after 1.1ps it only dropped the presence of either NH3 (400 ppm) or water vapor (10%)the removal to 67% of its peak and therefore the FWHM would be much longer than of both NO and NO, were enhanced [95]. When both water vapor and 1.1 ps. It is known that very narrow pulses ( 4 0 0 ns) are required ammonia were added simultaneously, the removal ratio increased sub- to obtain effective denitrification and desulfurization of the flue gases stantially For example at 30 W input power, without additives the [4,81]. The pulse voltage shown in Figure 3 of [26] to which the dc removal ratios of NO and NO, were, respectively 18%and 1% (Fig- performance of NO, removal was compared was actually that, as stated ure 54). For NO this increased to 6O%, 75% and 89% and for NO, to by the authors, of the charging voltage to the pulse forming network 20%, 33%and 74%, respectively, when 400 ppm of NH3,10% HzO, and and not the peak of the pulsed voltage. The charging voltage is usually 400 ppm NH3 plus 10% of HzO were added to the gas mixture (Fig- lower than the peak of the pulsed voltage provided by the dc charged ure 54) [95]. It was reported that there was no excess ammonia left in circuit. the exhaust gases due to the absorption of NH3 by H20 [95]. Figures 56 and 57 show the dependence of the removal ratios of NO,

Figure 54. NO (a) and NO, (b) removal ratios as a function of input of power using positive pulsed voltage in a mixture of 400 ppm NO, 10%02,10%0CO2 and 80% NZfor four conditions of with and without additions of N h and Hz0. Discharge reactor: coaxial glass dielectric barrier 1 mm thick, 23 mm inner diameter, 210 mm long, stainless steel wire 0.2 mm in diameter; flow rate 2 l/min; temperature 150C; pressure l.01x105 Pa. In (a) the top curve is for added 10%HsO t 400 ppm NH3 [95].

Figure 55. Reduction of NOx for different stoichiometric concentrations of the additives methane and ammonia using negative dc or pulsed vdltagesa Gas mixture; 438 ppm NO, 83.96% Nz, 8% 0 , 2 8% CQ; pressure 1.0ix105Pa; flow rate 5 l/min; discharge reactor: plane-plane; N h (or CHa)/(NO,), = 1 (1st); NH3 (or CHa)/(NO,),= 0.5 (0,5 st); (NO.& = 438 ppm NO; Voltage for pulse was charging voltage of pulsed circuit and not peak of pulsed voltage [26].

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7 N o . 5, October 2000

675

,
08

,
14

,
16

, . , I
18

10

12

20

MR

G
D .-

' 0

100

ti m
70& .

0'

-lean
o

electrode Contaminated electrode

,-. E
D

0 " ' Lo 6 0 -

,8

!g
a

50

SO.,

. ,

, . ,

. ,
PlQ=l.87 Wh"'
Ny-360 ppm

Figure 57. SO2 removal efficiency as a function of MR. Conditions are as in Figure 56 [25].

01 U

SO

100

150

200

250

300

676

Hackam et al.: Air Pollution Control by Electrical Discharges


same gas mixture, the same gas flow rate, without H20, the NO, removal steadily increased with increasing pulsed power with (200 to 800 ppm) and without addition of NH3 (Figure 60) [95]. Figure 49 shows that the decrease in the magnitude of the NO concentration using positive discharge in a glass coaxial dielectric barrier discharge was larger at 40 and 30 W power input to the discharge than at 20 W for all temperatures in the range 25 to 220C [9].
Pulse (Dry) Pulse (Semi-wet)

electron beam irradiation [5] the NO, removal also was enhanced with increasing the added ammonia concentration. The addition of double bond structure hydrocarbons (2-propane-l01) to a mixture of 400 ppm NO, 10% C02, 80% N2 and 10% 0 2 at a flow rate of 2 l/min (residencetime in the coaxial reactor 1.58 s) led to an enhanced removal of NO, [103]. A pulsed coaxial glass dielectric barrier reactor (0.5 mm diameter stainless steel wire, 13.3 mm inner diameter of the glass tube and 380 mm long) was employed at room temperature [103]. The addition of 2-propanol, which does not have a double bond did not result in a significant reduction of NO, [103]. Using a copper coated zeolite catalyst (Cu-ZSM-5) with hydrocarbons promoted an improvement in the removal of NO, using positive pulsed voltage in a glass dielectric discharge reactor [85,103].

3
. 3

80

i i
Q)

18 POLLUTANT REMOVAL AND INPUT POWER

40 -

l8kHz (Semi-wet)

A number of studies have been reported on the dependence of the removal ratio of the pollutant on the input power to the discharge [9, 64,85,93,95,96,103]. Discharge power [W] Figure 26 shows the dependence of the NO, removal ratio on the inFigure 61. Removal efficiency of N h from air as a function of input put power to the discharge in the form of an arc plasma in a simulated discharge power using positive square pulses, 60 Hz and 18 kHz. Dismixture of gases of 346 to 2317ppm NO, to 11%02, to 8%CO:! and balcharge reactor: coaxial glass dielectric barrier, 0.2 mm diameter wire, ance N2, including an unknown quantity of water vapor [64]. Figure 26 18 mm inner diameter of glass tube, 1 m thick, 150 mm long; semim wet: filtered paper soaked with CaCh solution and placed on inside shows that the NO, removal ratio steadily increased with increasing surface of glass tube to absorb and retain moisture from air; gas resipower and reached 100%when the plasma power was in the range 0.8 dence time 0.6 s; initial NH3 concentration 18 to 25 ppm [93]. to 1.2 kW. At high power, the temperature of the plasma increased and more NO was generated which resulted in lowering the NO, removal Figure 61 shows the removal efficiency of NO, from air as a function efficiency [64]. This occurred for all inlet values of NO [64]. of input power to the discharge using positive square pulses, 60 Hz and 18 kHz in a coaxial glass dielectric barrier discharge [93]. It will be observed that the pulse voltage performed better than the ac for dry air. This was attributed to more intense and uniform plasma obtained by the positive pulsed voltage [93]. For the semi wet condition (filter paper was soaked with CaC12 solution and placed on the inside surface of the glass tube to absorb moisture from the air and retain it) the maximum removal performance was at about the same level (Figure 61) [93]. In all cases the removal ratios of NO, increased with increasing discharge power (Figure 61) [93]. NH3 is dissociated by the discharge and reacts with H20 to form NH4N03 [78,106,107]. This explains the finding in Figure 61 that for the semi-wet condition the removal efficiency of NO, 0 IO 20 30 40 Input power [W] was higher than for dry air.
Figure 60. Dependence of NO, removal on input power of pulsed voltage into a coaxial glass dielectric barrier discharge reactor. Gas mixture, 400 ppm NO, 10% 02, 10% COz, 80% N2; discharge reactor, gas pressure and flow rate as in Figure 54; dry reactor, 25C 1951.

19 ENERGY DENSITY REQUIREMENTS

It is of more general interest to ascertain the energy density required Figure 54 shows the dependence of the removal ratio of NO and to remove the pollutants from the flue gases than the power fed into the NO, on the input power of positive pulsed voltage into a coaxial glass discharge. This is for the purposes of comparison of different discharge discharge reactor in 400 ppm NO, 10% 0:!,10% C02, and 80% NZ at techniques employed as well as with the alternative method of electron 150C [95]. The removal ratios of both NO and NO, steadily increased beam irradiation and for determining the energy cost associated with with increasing input power and reached constant removal ratios, de- the removal process. The selection of the technology for implementing pending on the additives used, in the range 20 to 30 W [95]. With fur- the pollution control will depend largely on the energy cost. ther increase in the input power to the discharge, a reduction in the Figure 62 shows the input energy density for removal of NO in a removal of both ensued (Figure 54). However, at 25C and using the mixture of 100 NO in nitrogen at l.01x105 Pa and 25C using pulse

IEEE Transactions on Dielectrics and Electrical Insulation


100

Vol. 7 No. 5, October 2000

677

80
Y

a .
60

? !
C

0"

! 40
b
0

0
220

b I

A beam

electron

b..

'..e
7..

100 ppm CCI, in Dry Air

25%
0 1 . ' " '
. . . I I .

. .

50

100

150

200

Input Energy Density (J/I)


Figure 62. Energy density requirements to remove NO in NZ using positive pulsed corona and electron beam methods. Cas mixture: 100 ppm NO; pressure 1.01x105Pa; temperature 25T; discharge reactor: coaxial, 1.5 mm wire diameter, 60 mm diameter metal tube, 300 mm long; pulse width 100 ns; electron beam system, 125 keV via titanium window 17.8pm thick [108]. corona discharge and electron beam reactors [108]. The discharge reactor employed a 1.5 mm diameter wire in a 60 mm diameter metal tube 300 mm long. The pulsed power was obtained from a magnetic pulse compressor and had 100ns FWHM. The electronbeam reactor employed an electron beam of 125 keV which was produced from a low pressure helium plasma and fed into the reactor containing the pollutant via 17.8 pm thick titanium window [108]. Figure 62 shows that the energy density required to remove NO completely was 120 J/1(0.134mol (NO)/kWh = 4.02 g (NO)/kWh) and 20 J/l (0.804 mol (NO)/kWh = 24.1 g (NO/kWh), respectively, using the pulsed corona and the electron beam methods in nitrogen. Therefore the removal of NO in dry nitrogen using pulsed corona is much less energy efficient compared to the electron beam process. The energy density requirements for reducing carbon tetrachloride (CCl4)in dry air (20% 02,8O%N2) at 25C are shown in Figure 63 [108]. It was suggested that the energy required for CC4 removal was determined by the creation of electron-ion pairs from ionization of nitrogen and oxygen in the gas mixture which is followed by dissociative attachment of CCll (e t CCla + Cl- t CC13) [108]. The energy density required for the reduction of methylene chloride (CH2Cl2)in dry air using pulsed corona is shown in Figure 64 for different gas temperatures in the range 25 to 300C [108]. It will be observed that the energy efficiency for the removal of methylene chloride in air increases substantially with increasing gas temperatures. In a mixture of 200 ppm NO, 5% 02,4% H20, 91% NPthe energy and requirement to remove NO using positive pulsed voltage in a coaxial discharge reactor is shown in Figure 42 [43]. It will be observed that a higher energy efficiency (mol NO/kWh) was found with decreasing pulse width from 120 to 40 ns. Typically, at 95% removal ratio of NO from the mixture, the energy efficiency was 0.76 and 0.3 mol/kWh (1

Input Energy Density (JA)


Figure 63. Reduction of carbon tetrachloride (CCla) in dry air vs. energy density using pulsed corona and electron beam methods. Gas mixture: 100 ppm CCla; 20% 02,80% Nz; other conditions as in Figure 62 [108].

Pulsed Corona Processing 160 ppm CH,CI, in Dry Air

3 1
0

'4

.
, . . .
/ . . . . I

'&--- . . . .

---- - - - A # . . . .

Figure 64. Reduction of methylene chloride (CHzC12)in dry air vs. input energy density using pulsed corona discharge for different gas temperatures. Gas mixture: 160 ppm CHzC12,20%02,80% Nz; other conditions as in Figure 62 [108]. mol NO = 30 g NO), respectively for 40 and 120 ns (Figure42) [43]. The removal energy efficiency depends strongly on the removal ratio of NO and it increaseswith decreasing removal ratio. For example, using 40 ns positive pulse width, the removal energy efficiency increased from 0.76 mol NO/kWh (22.8 g NO/kWh at 95% NO removal ratio to 2.1 mol NO/kWh (63 g NO/kWh) at 20% removal ratio [43]. A possible reason for the reduction of the removal energy efficiency (mol/kWh) with increasing removal ratio of NO shown in Figure 42 is because the increase in the removal ratio was obtained with increased energy input into the discharge (higher pulse rate in Figure 42) while

678

50 PPm 100 ppm A 200 ppm 400 ppm ' 600 ppm I 0 800 ppm

Hackam et al.: Air Pollution Control by Electrical Discharges

5
A

L:,I,:
8

Pulse repetition rate [pps]


50ppm 100ppm

+
' I

.
(c

10

A 200ppm

400ppm 6OOppm a00ppm


35

30

.......-..+... : , ,,> --~

in-phase ~ m i n :* ..

25

' I

Figure 65. NO and NO, removal energy efficiency as a function of pulse repetition rate for different initial concentration of NO. Gas mixture: 5% 02,4% H20, 91% N2, 50 to 500 ppm NO; pressure l.01x105Pa; temperature 25C; flow rate 2 limin; (reduced to O ) T; discharge reactor: 0.5 mm stainless steel diameter wire, 76 mm inner diameter copper cylinder, 800 m m long; positive pulse, peak voltage 53 kV, peak current 54.2 A, pulse width 100 ns [109].

z f
P

20

.'I

,' ,

, '

, .: '

,<: ,

out-af-phase

- --_
-9

PUmin

.t

aut-of-ptiase
1Umin

the amount of NO available for destruction was steadily decreasing. in-phase lUmin 5 This can be seen from Figure 65 where the energy efficiency decreases at a fixed initial concentration of NO with increasing pulse rate, and also decreases at a fixed pulse rate with decreasing initial concentration 0 . of NO [109]. Figure 65 shows that the removal efficiency of both NO 0 0.2 0.4 0.6 0.8 1 1.2 1.4 [HO]o/[NH3]o and NO, increase with increasinr initial concentration of NO from 50 to 600 ppm and then decrease withy further increase in NO [109]. Similar Figure 66. Dependence of energy yield (g/kWh) of NO, reduction behavior was reported for the removal efficiency of NO, (Figure 65) on the injected ammonia using combined dielectric and surface dis[109]. charges. Discharge reactor is shown in Figure 4. gas mixture: N2, COz, 0 (proportion not specified),800 ppm NO; flow rates 1 and 2 l/min; 2 Masuda and Nakao reported that the removal energy efficiency of pressure 1.01~10~1281. Pa NO in air depended on the power input, the gas flow rate and the temperature [all. For a mixture of 180 ppm NO in dry air with 180 ppm Figure 66 shows the energy yield (g/kWh) for NO, reduction usNH3 and using coaxial discharge reactor of 3 mm wire in diameter, 90 mm in diameter metal cylinder and 1.3m long the energy yield of ing ac (60 Hz) in a combined dielectric barrier and surface discharges NO removal was 20 g/kWh [NI. This was higher than that of 3 to reactor with addition of ammonia [28]. It will be observed that the energy yield increased with decreasing ammonia concentration and with 5 g/kWh obtained earlier by the same group using dc [U]. increasing flow rate. Further when the ac voltage supplied to the dielecUsing pulsed power and a coaxial discharge reactor Dinelli et al. reported a yield of 25 g NO/kWh for the removal of 50% NO, from the tric barrier and the surface discharges were out-of-phase a larger yield flue gases of a coal-burning power station when ammonia was used was attained for the same amount of injected ammonia [28]. Figure 66 and the initial NO, concentration was in the range 500 to 550 ppm [97]. shows that the highest yield of 32 g NOJkWh was obtained at 2 l/min Shimizu et al. reported that using a positive pulse to treat the ex- gas flow rate at an initial concentration of 800 ppm NO and injected haust gases from diesel engine which were at 150C required 220 J/g ratio of NO/NH3 of 1.2 [28]. For flue gases from a boiler of a thermoelectric plant and using a to attain 60% removal of NO,for dry reactor, and 60 J/g (70% of NO, removed) for wet reactor [96]. The initial concentrations of NO were coaxial discharge reactor, an energy density input to the discharge of 100 to 120 ppm and of NO, 110 to 160 ppm. Other constituents in the 12 Wh/m3 was required to attain a complete removal of SO2 for initial

i-------'-; 1
,? .

IEEE Transactions on Dielectrics and EJectrical Insulation


concentrations from 443 to 645 ppm of NO, with ammonia injection [98]. The concentration of ammonia at the outlet of the discharge reactor (called ammonia slip) was generally t4 ppm [98]. For an inlet concentration of NO, of 250 ppm and 60% removal ratio 12 to 15 Wh/m3 was supplied to the pulsed discharge [98]. Vogtlin and Penetrante reported on the required energy density into the discharge for the removal of NO and SO, in air with a large injection of n-octane (n-octane: NO, = 8: 1)using pulse power [110]. At 224"c, 750 ppm of the initial 800 ppm of S O were removed with an energy density input of 9.5 Whim3, and 500 ppm was removed when the initial concentration was 1100 ppm [110]. This corresponded to 18 eV per removed NO molecule (1 eV = 3.216 kJ/g NO) and 29 eV per removed SO, molecule. The influence of the capacitance of the pulse forming circuit on the energy density consumed for the removal of NO and SO2 from simulated flue gases (air containing 3.5% H20) and initial concentrations of 200 ppm NO and 200 ppm SO2 was investigated [lll]. Generally lower energy density was required to remove S O and SO2 with smaller capacitance. The energy density required for the removal of various volatile compounds using pulsed corona in dry air are shown in Figures 33 to 35 [72]. The energy density u/l) required to remove trichloroethylene (TCE) in dry air was the same using pulsed dielectric barrier, ac dielectric barrier and flat-bed (figure not shown) reactors. The energy density increased with decreasing the removal ratio of TCE [72]. The removal efficiency of methylene chloride (MC) TCE from dry and air were found to increase with increasing energy density u/1) into the discharge using a ceramic dielectric barrier with pulsed corona [112]. 1400 J/l was required to remove 95% of TCE and 78% of MC [112]. Takaki et RI. reported that removing 50% of NO from a mixture of 200 ppm NO, 90% Nz, and 10%0 using a dielectric barrier flat glass 2 discharge reactor with ac (10 to 100 kHz) required 18 g NO/kWh [70]. Figure 31 shows the reduced NO and SO, in this mixture vs. energy density O/l) [70]. The removal efficiency of NO? from flue gases of a coal-burning plant us. the specific energy input into the discharge of pulsed corona reactor is shown in Figure 67 [99]. The removal ratio steadily increased with increasing energy density (Wh/m3)into the discharge. The reactor was in the form of a duct with multiple wires 3 mm in diameter. At a removal ratio of 50% of NO2 (initial concentration of NO, = 515 ppm), 25 g NO/kWh was consumed 1991. The SO2 removal ratio as a function of energy density u/g) in air without addition of NH3 is shown in Figure 68 for a mixture of 1000 ppm S02,6% H20, and using multi pin needles-plane geometry reactor with a negative dc voltage [102]. It will be observed that the removal efficiency increased linearly with increasing energy density. The addition of water was found to be beneficial for the removal of SO2. For example, when a fixed amount of 20 Jig was deposited into the discharge, 75 ppm of SO2 were removed (from an initial concentration of 300 ppm) at a water content of 1.5%,and 110 ppm were removed (initial concentration of 1000ppm) at 6% water content [102]. It was claimed [lo21that the energy requirement using the multi pin needle to plane geometry with negative dc voltage at high current density of 10 mA/cm2 resulted in 8 J/g per (100 ppm SO2) was comparable
N

Vol. 7 N o . 5, October 2000

679

-1

5 50
40

30

10

12 14

16 18 20

Specific energy w / I / N ~ ~ )

Figure 67. NO2 removal ratio from flue gases of coal burning plant us. energy density (Wh/m3) input into the discharge. NO, 515 ppm; SO2 360 to 550 ppm; discharge reactor: duct with multiple wires 3 mm in diameter; solid particulates 150 mg/m3; flow rate 600 m3/h; temperature 70 to 100C; residence time 4 to 7 s; NHJ/(NO, t SO2) = 0.7 to 0.8; pressure 1 . 0 1 ~ 1 Pa~1991, 0

Figure 68. SO2 removal ratio from air vs. energy density u/g) input into the discharge without addition of ammonia. Other conditions are as in Figure 52 [102].

to the best results for the electron beam technology [113] and pulsed techniaues 1971.
I

Using 100 kV, 10 ns rise time, 200 ns wide pulses, it was possible to decompose completely in air styrene, toluene, ethylene and S O with a coaxial reactor having a wire 0.25 to 1mm in diameter, 250 mm in diameter stainless steel cylinder, 3.5 m long [87]. The specific energy required and the initial concentrations were, respectively for NO: 15 kWh/kg, 213 ppm; styrene: 7 kWh/kg, 30 to 190 ppm and ethylene: 12 kWh/kg, 150 to 2500 ppm [87]. Pentane, 1,1,1 trichloroethylene (TCE), propane and butane were not completely decomposed in air. The required energy density was higher and varied from 24 kWh/kg for toluene to 180 kWh/kg for both 1,1,1 TCE and propane [87]. Figures 69 and 70 show the dependence of SO2 and NO reduction on

680

Hackam et al.: Air Pollution Control by Electrical Discharges


Figure 69 shows that negative dc corona is superior than positive dc for reducing NO and NO, [104]. A comparison between Figures 70 and 69 shows that the energy density input necessary to reduce NO, NO, and SO1 is much less for pulsed than for dc corona. To obtain 90% reduction of NO, from an initial concentration of 300 ppm, 50 J/g was required using pulsed corona, while for dc to obtain 70% reduction 600 J/g was consumed (Figure 69) [104].

20 CONCLUSIONS
HE current state of the art in using discharges to reduce pollutants present in air that originate from a variety of sources including coalburning power plants, diesel engine exhausts, industrial plants, as well as simulated mixtures of gases used in laboratories, is reviewed. It appears that non-thermal discharges using very fast rise ( 4 0 ns) and short duration pulses of <50 ns are likely to offer promising prospects. This is because the energy supplied into the discharge is predominantly used to create energetic electrons which produce radicals such as N, OH and 0 that lead to the removal of the pollutants. The energy required to remove the pollutants is of prime concern. It would serve as one of the main considerations in the selection of the technology to be used ultimately Substantial effort is being expended worldwide to design more efficient pulsed power sources to be employed for pollutant control as well as improved discharge reactor geometries. The requirements of recent national legislation and international agreements to reduce significantly the amounts of emissions of pollutants into the air necessitate a development of a cost effective system that can be used in electrical generation and other plants. This has become of some urgency of late. It is expected that the research into new techniques to remove yollutants, better understanding of the chemical and physical processes involved as well as development and testing prototypes of discharge reactors in power and other plants will continue at an accelerated rate.

Figure 69. Dependence of SO2 a n d NO, reduction o n energy input into positive and negative d c discharge with injected ammonia.
Gas mixture: 70.4% N2, 1.8% 0 , 2 9.9% CO2, 17.9% H20; temperature 100C; gas flow rate 0.111 g/s; discharge reactor: 3 stainless steel wires 0.3 m m in diameter to plate; gap distance 2 cm; gas residence time 12.3 s; gas flow speed 4.06 cm/s; injected NH3 diluted in NZ (1:lO); NO, 300 ppm; SO2 300 ppm [104].

REFERENCES
a 0

[l] S. Tsukamoto, T Namihira, D. Wang, S. Katsuki, H. Akiyama, M. Koike, Y.Uchida, . E. Nakashima and A. Sato", NO, and SO2 Removals By Pulsed Power at a Thermal

4L 0-

20

40
Spcilic diichmge input (119)

60

' I 80

Figure 70. As for Figure 69 except for positive and negative pulsed voltage. Pulsed width 100 ns [104].

the energy density u/g) in a mixture of 70.5% Nz,1.8%02, 9.9% COz, 17.9%H20 and ammonia for dc and pulsed voltages, respectively [50, 1041. Both voltage polarities were employed. The ammonia was first 2 diluted with N at a ratio of 1:lO and the amount injected into the gas mixture was equivalent to convert SO2 and NO to ammonium sulfate and ammonium nitrate, respectively

Power Plant", 12th IEEE Int. Pulsed Power Conf., pp, 1330-1333,1999, [2] B. M. Penetrante, "Removal of NOx From Diesel Generator Exhaust by Pulsed Electron Beams", 11th IEEE Int. Pulsed Power Conf., pp, 91-96,1997. [3] G. Dinelli, L. Civitano and M. Rea, "Industrial Experiments on Pulsed Corona Simultaneous Removal of NO, and SO2 from Flue Gas", IEEE Trans. Indust. Appl., Vol. 26, pp. 535-541,1990. [4] S. Masuda, "Pulse Corona Induced Plasma Chemical Process: a Horizon of New Plasma Chemical Technologies", Pure and Appl. Chem., Vol. 60, pp, 727-731,1988. [5] J. S. Chang, I C. Looy, K. Nagai, T.Yoshioka, S. Aoki and A. Maezawa, "Preliminary ? Pilot Plant Tests of a Corona Discharge-ElectronBeam Hybrid Combustion Flue Gas Cleaning System", IEEE Trans. Ind. Appl., Vol. 32, pp. 131-137,1996. [6] Y. Mok and I-S. Nam, "Positive Pulsed Corona Discharge Process for Simultaneous Removal of SOz and NO, from Iron-Ore Sintering Flue Gas", IEEE Trans. Plasma Sci., Vol. 27, pp. 1188-1196,1999. [7] A. Mizuno, J. S. Clements and R. H. Davis, "A Method for the Removal of Sulfur Dioxide From Exhaust Gas Utilizing Pulsed Streamer Corona for Electron Energization", IEEE Trans. Ind. Appl., Vol. 22, pp. 516-522,1986. [8] D.Bhasavanich, S. Ashby, C. Deeney and L. Schlitt, "Flue Gas Irradiation Using Pulsed Corona and Pulsed Electron Beam Technology", 9th IEEE Int. Pulsed Power Conf., pp, 441-444,1993.

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 7 N o . 5, October 2000

681

[30] H. Matzing, "Chemical Kinetics Model of S02/NO, Removal by Electron Beam",

in Non-Tliermnl Tediniqrres for Pollution Conlrol, Eds. B. M. Penetrante and S. E. Schultheis, Part A, pp. 59-64,1993. [31] R. l? Dahiya, S. K. Mishra and A. Veefkind, "Plasma Chemical Investigations for NO, and SO2 Removal from Flue Gases", IEEE Trans. Plasma Sci., Vol. 21, pp. 346348,1993. [32] H. Akiyama, K. Kawamura, T. Takeshita, S.Katsuki, S. Maeda, S. Tsukamoto and

M.Murata, "Removal of NO, Using Discharges by Pulsed Power", 10th IEEE Int.
Pulsed Power Conf., pp, 133-137,1995. [33] I. Sardja and S. K. Dhali, "Plasma Oxidation of SO2", Appl. Phys. Lett., Vol. 56, pp, 21-23,1990 [34] J. Li, W. Sun, B. Pashaie and S. K. Dhali, "Streamer Discharge Simulation in Flue Gas", IEEE Trans. Plasma Sci., Vol. 23, pp. 672-678,1995. [35] J. J, Lowke and R. Morrow, "Theoretical Analysis of Removal of Oxides of Sulfur and Nitrogen in Pulsed Operation of Electrostatic Precipitators", IEEE Trans. Plasma Sei., Vol. 23, pp, 661-671,1995. [36] B. M. Penetrante, M. C. Hsiao, B. T. Merritt, G. E. Vogtlin and l? H. Wallman, "Comparison of Electrical Discharge Techniques for Non-Thermal Plasma processing of NO and NO2", IEEE Plasma Sci., Vol. 23, pp, 679-687,1995. [37] U. Kogelschatz, "Advanced Ozone generation", in Process Technologies for Wnter Trenlment, Ed. S. Stucki, Plenum, NY, pp. 87-120,1988, [38] U. Kogelschatz, B. Eliasson and M. Hirth, "Ozone Generation from Oxygen and Air: Discharge Physics and Reaction mechanisms", Ozone Science and Engineering, Vol. 9, pp, 367-377,1987. [39] S. Masuda, "Destruction of Gaseous Pollutants and Air Toxics by Surface Discharge Induced Plasma Chemical Process (SPCP)and Pulse Corona Induced Plasma Chemical process (PPCP)", in Non-Tliermnl Plnsnio Tecliniqucs for Pollstion Control, Eds. B. M. Penetrante and S. E. Schultheis, Part B, pp. 199-210,1993, Ohkubo, S.Kanazawa, Y. Nomoto, J. S. Chang and T. Adachi, "NO, Removal by [40] T, a Pipe With Nozzle-Plate Electrode Corona Discharge System", IEEE Trans. Indust. Appl., Vol. 30, pp. 856-861,1994. [ U ] T. Ohkubo, S. Kanazawa, Y.Nomoto, J, S. Chang and T.Adachi, "Time Dependence of NO, Removal Rate by a Corona Radical Shower System", IEEE Trans. Indust. Appl., Vol. 32, pp, 1058-1062,1996, 1421 H. Akiyama, Y.Nishihashi, S. Tsukamoto, T. Sueda and S. Katsuki, "Streamer Discharges by Pulsed Power on a Spiral Transmission Line", 11th IEEE Int. Pulsed Power Conf., pp. 109-114,1997, Namihira, S. Tsukamoto, D. Wang, S.Katsuki, R. Hackam, H. Akiyama, U. Uchida 1431 T. and M. Koike, "Improvement of NO, Removal Efficiency Using Short Width Pulsed Power", IEEE Trans. Plasma Sci., Vol. 28, April issue, 2000. [44] W. J. M. Samaranayake, Y.Miyahara, T. Namihira, S.Katsuki, T. Sukugawa, R. Hackam and H. Akiyama, "Pulsed Sreamer Discharges Characteristics of Ozone production in Dry Air", IEEE Trans. Diel. Elect. Insul., Vol. 7, pp, 254-260,2000. [45] E Hegler and H. Akiyama, "Ozone Generation by Positive and Negative Wire-toPlate Streamer Discharges", Japan J, Appl. Phys., Vol. 36, pp, 53355339.1997. [46] B. Held, "Coronas and Their Applications", 11th Int. Conf. of Gas Discharges and Their Applications, Tokyo, Vol. 2, pp, 514526,1995, f [47] T. Namihira, Trentnrent o Exhnust Gnses by Pulsed Poser, Master Thesis, University of Kumamoto, 1999 (In Japanese) 1481 D. Wang, NO, Renlovol Using Pulsed Power, Master Thesis, University of Kumamoto, 2000 (In Japanese) [49] A. Jaworek, A. Krupa and T. Czech, "Decomposition of NO2 in Oxygen-Free NOz: N2 Gas Mixture by Back-Corona Generated Plasma", Plasma Phys., Vol. 36, pp, 619629,1996.

1501 K. Onda, Y.Kasuga, K. Kato, M. Fujiwara and M. Tanimoto, "Electric Discharge Removal of SO2 and NO, from Combustion Flue Gas by Pulsed Corona Discharge", Energy Convers. Mgnt., Vol. 38, pp. 1377-1387,1997. [51] A. Jaworek, J. Mizeraczyk, A. Krupa and T. Czech, "Removal of NOx from NOz: N O N2 Mixture by a Pulsed or Dc Streamer Corona in a Needle-to-Plate Reactor", Czechslovak J, Phys., Vol. 45, pp. 1049-1061,1995, 1521 J. S.Chang and I. Maezono, "The Effect of Electrohydrodynamic Flow on a Corona Torch", J. Electrost., Vol. 23, pp. 323-330,1989, [53] J, S. Chang, "The Role of H20 on the Formation of N b N O 3 Aerosol Particles and De-NO, Under the Corona Discharge Treatment of Combustion Flue Gses", J. Aerosol Sci., Vol. 20, pp, 1087-1080,1989,

682

Hackam et al.: Air Pollution Control by Electrical Discharges


1761 B. Eliasson, E G. Simon and W. Egli, "Hydrogeneration of SO2 in a Silent Discharge", in Non-ThermalPlnsiiro Tecliniquesfor Pollution Control, Eds. B. M.Penetrante and S. E. Schultheis, Part B, Springer-Verlag, pp. 321-338,1993. [77] B. Eliasson and U. Kogelschatz, "Modeling and Applications of Silent Discharge Plasmas", IEEE Trans. Plasma Sci., Vol. 19, pp, 309-321,1991. [78] J. S. Chang, E A. Lawless and T. Yamamoto, "Corona Discharge Processes", IEEE Trans. Plasma Sci., Vol. 19, pp. 1152-1166, 1991. 1791 J, S. Clements, A. Mizuno, W. C. Finney and R. H. Davis, "Combined Removal of SOL, NO, and Fly Ash from Simulated Flue Gas Using Pulsed Streamer Corona", lEEE Trans. Indust. Appl., Vol. 25, pp. 62-69,1989. [80] H. Akiyama, "Pollution Control by Pulsed Power", Proc. Int. Power Electronics Conf. Yokohama, pp. 1397-1400,1995, [81] S. Masuda and H. Nakao, "Control of NOx by Positive and Negative Pulsed Corona Discharges", IEEE Trans. Indust. Appl., Vol. 26, pp, 374383,1990. [82] I. Gallimberti, "Impulse Corona Simulation for Flue Gas Treatment", Pure and Appl. Chem., Vol. 60, pp, 663-674,1988. [83] Q. Li, C. Ning, S. Li and J, Li, "Simultaneous Removal of SO2, NO, and Fly Ash from Flue gas Utilizing HV Pulsed Corona Discharge", 7th Asian Cod. On Electrical Discharges, pp. 173-177,1994, 1841 2. Mmin, W. Yan and W. Rongyi, "Optimization of a Generator/Reactor System for S02/NOx Removal by Narrow Pulse HV Corona Discharge", 7th Asian Conf. On Electrical Discharge, pp. 182-187,1994. [85] T. Oda, T. Kato, T. Takahashi and K. Shimizu, "Nitric Oxide Decomposition in Air by Using Non-Thermal Plasma Processing with Additives and Catalyst", J, Electros., Vol. 42, pp. 151-157,1997, [86] R. Wang, B. Zhang, B. Sun, N. Wang, Y. Wu, W. Gong and S. Liu, "Apparent Energy Yield of a High Efficiency Pulse Generator with Respect to SO1 and NO, Removal", J. Electrost., Vol. 34, pp. 355-366,1995. I871 E. J. M. vanHeesch, H. W. M. Smulders, S. V. M. vanpaasen, I? P M. Blom, E M. van Gompel, A. J. l? M. Staring and K. J. Ptasinski, "Pulsed Corona for Gas and Water Treatment", 11th Int. Pulsed Power Conf., pp, 103-108,1997, [88] M. G. Grothaus, R. K. Hutcherson, R. A. Korzekwa and R. Roush, "Coaxial Pulsed Corona Reactor for Treatment of Hazardous Gases", 9th IEEE Pulsed Power Conf., pp. 180-183,1993, [89] M. G. Grothaus, R. K. Hutcherson, R. A. Korzekwa, R. Brown, M. W. Ingram, R. Roush, S. E. Beck, M. George, R. Pearce and R. G. Ridgeway, "Gaseous Effluent Treatment Using a Pulsed Corona Discharge", 10th IEEE Int. Pulsed Power Conf., pp, 124-132,1995. [90] A. Mizuno, A. Chakrabarti and K. Okazaki, "Application of Corona Technology in the Reduction of Greenhouse Gases and Other Gaseous Pollutants", in Non-Tkermal Plasma Techniquesfor Polliition Control, Eds., B. M. Pentrante and S. E. Schultheis, Part B, Springer-Verlag, pp. 165-186,1993. [91] D. J. Helfritch, "Pulsed Corona Discharge for Hydrogen Sulfide Decomposition", in Non-Tliermnl Pinsinn Techniquesfor Pollution Control, Eds. M. B. Pentrante and S. E. Schultheis, Part 8, Springer-Verlag, pp. 211-222,1993, [92] G. J, Roth and M. A. Gunderson, "Laser-induced Fluorescence Images of NO Distribution After Needle-Plane Pulsed Negative Corona Discharge", IEEE Trans. Plasma Sci., Vol. 27, pp, 28-29,1999. [93] A. Chakrabarti, A. Mizuno, K. Matsouka and S. Furuta, "Gas Cleaning with SemiWet Type Plasma Reactor", IEEE Trans. Indust. Appl., Vol. 31, pp. 500-506,1995, [94] R. A. Roush, R. K. Hutcherson and M. W. Ingram, "Effects of Pulse Risetime and Pulsewidth on the Destruction of Toluene and NO, in a Coaxial Pulsed Corona Reactor'', 22nd Int. Power Modulator Symposium, Boca Raton, USA, pp, 79-84,1996. [95] A. Mizuno, K. Shimizu, T. Matsuoka and S. Furuta, "Reactive Absorption of NO, Using Wet Discharge Plasma Reactor", IEEE Trans. Indust. Appl., Vol. 31, pp, 14631468,1995. [96] K. Shimizu, K. Kinoshita, K. Yanagihara, R. S. Rajanikanth, S. Shinji and A. Mizuno, "Pulsed -Plasma Treatment of Polluted Gas Using Wet-/Low Temperature Corona Reactors", IEEE Trans. Indust. AppI., Vol. 33, pp, 1373-1380,1997, [97] G. Dinelli, L. Civitano and M. Rea, "Industrial Experiments on Pulse Corona Simultaneous Removal of NO, and SO2 from Flue Gas", IEEE Trans. Indust. Appl., Vol. 26, pp. 535-541,1990, [98] L. Civitano, "Industrial Applications of Pulsed Corona Processing to Flue Gas", in Non-Tliermnl Plasma Techniquesfor Pollution Control, Eds. B. M. Penetrante and S. E. Schultheis, Part 8, Springer-Verlag, pp, 103-130,1993,

1541 S. Kanarawa, J. S. Chang, G. Round, G. Sheng, T. Ohkubo, Y. Nomoto andT. Adachi, "Removal of NO, from Flue Gas by Corona Discharge Activated Methane Radical Showers", J. Electrostatics, Vol. 40 & 41, pp, 651-656, 1997. [55] S. Kanazawa, J. S. Chang, G. F. Round, G. Sheng, T. Ohkubo, Y. Nomoto and T. Adachi, "Reduction of NOx from Flue Gas by Corona Discharge Activated Ammonia Radical Showers", J, Combust. Sci. Technol., Vol. 133, pp. 93-105,1998. [56] H. Matzing, "Chemical Kinetics of Flue Gas Cleaning by Irradiation With Electrons", Adv. Chem. Phys., Vol. 80, pp, 315402,1991 [57] I. Maezono and J. S. Chang, "Reduction of CO2 from Combustion Gases by DC Corona Torches", IEEE Trans. Indust. Appl., Vol. 26, pp. 651-655,1990, [58] I Maezono and J. S. Chang, "Flow Enhanced Corona Discharge Torch, J. AppI. Phys., Vol. 59, pp, 2322-2324,1988. [59] R. A. Hampson, "Chemical Kinetic and Photo Chemical Data Sheet for Atmospheric Reactions", US Department of Transportation, Report FAA-EE-80-17,1980. [60] T. Morimme and S. Nedachi, "Exhaust Cleaning by Plasma Jet", Int. J. JSME, Ser. 11, Vol. 33, pp. 548-554,1990, [61] T. Morimune and S. Nedachi, "Removal of Pollutants Contained in Combustion Flue Gas by Plasma Injection", Trans. JSME, Ser. B, Vol. 56, pp. 3155-3159,1990. [62] H. Behbahani, A. M. Warris and E J. Weinberg, "The Destruction of Nitric Oxide E by Nitrogen Atoms from Plasma Jet", Combust. Sci. Technol., Vol. 30, pp. 289-302, 1983. [63] T. Morimune and Y. Ejiri, "Removal of NOx Contained incombustion Exhaust Gas by N2 Plasma Injection", Trans. JSME, Ser. B, Vol. 58, pp. 25842594,1992. [64] T. Morimune, Y Ejira and T. Tsukakoshi, "Removal of NO, from Exhaust Gas by N2 Arc Plasma Injection", Int. J. of Experimental Heat Transfer, Thermodynamics and Fluid Mechanics (Experimental Thermal and Fluid Sci.), Vol. 8, pp, 175-180,1994, [65] T.Morimune and Y. Ejira, "Removal of Nitrogen Oxides Contained in Combustion Exhaust Gas by Nitrogen Plasma Injection", JSME Int. J. Ser. B, Vol. 37, pp. 945-950, 1994. [66] A. Ogata, N. Shintani, K. Mizuno, S. Kushiyama and T. Yamamoto, "Decomposition of Benzene Using a Non-Thermal Reactor Packed with Ferroelectric Pellets", IEEE Trans. Indust. Appl., Vol. 35, pp. 753-759,1999. [67] M. B. Chang, M. J. Kushner and M. J. Rood, "Gas Phase Removal of NO From Gas Streams Via Dielectric Barrier Discharges", Environ. Sci. Technol., Vol. 26, pp. 777781,1992. [68] S. Futamura, A. Zhang and T. Yamamoto, "Mechanisms for Formation of Inorganic Byproducts in Plasma Chemical Processing of Hazardous Air Pollutants", IEEE Trans Indust. Appl., Vol. 35, pp. 760-766,1999. 1691 S. K. Dhali and I. Sardja, "Dielectric Barrier Discharge for Processing of S02/NOX", J. Appl. Phys., Vol. 69, pp, 6319-6324,1991. [70] K. Takaki, M. A. Jani and T. Fujiwara, "Removal of Nitric Oxide in Flue Gases by Multipoint to Plane Dielectric Barrier Discharge", IEEE Trans. Plasma Sci., Vol. 27, pp. 1137-1145,1999, [71] K. Fujii, M.Higashi and N. Suzuki, "Simulation Removal of NO,, CO,, SO, and Soot in Diesel Engine Exhaust", in Non-Tlierinal Plasma Techniqiies for Pollution Control, Eds. B. M.Penetrante and S. E. Schultheis, Part B, Springer-Verlag, pp. 257-279, 1993. [72] R. Korrekwa, L. Rosocha and Z. Falkenstein, "Experimental Results Comparing Pulsed Corona and Dielectric Barrier Discharges for Pollution Control", 11th IEEE Int. Pulsed Power Conf., pp. 97-102,1997, [73] T. Yamamoto, P A. Lawless, M. K. Owen, D. S. Ensor and C. Boss, "Decomposition of Volatile Organic Compounds by a Packed-Bed Reactor and a Pulsed-Corona", in Non-Tkermal Plasm Techniques for Polliitioii Control, Eds. B. M. Penetrante and S. E. Schultheis, Part B, Springer-Verlag, pp. 223-238,1993. [74] L. A. Rosocha, G. K. Anderson, L. A. Bechtold, J. J. Coogan, H. G. Heck, M. Kang, W. H. McCulla, R. A. Tennant and P J. Wantuck, "Treatment of Hazardous Organic Wastes Using Silent Discharge Plasmas", in Non-Thermal Plasm Techniquesfor Pollution Control, Eds. B. M. Penetrante and S. E. Schultheis, Part 8, Springer-Verlag, pp. 281-308,1993, [75] W. C. Neely, E. I. Newhouse, E. J. Clothiaux and C. A. Gross, "Decomposition of Complex Molecules Using Silent Discharge Plasma processing", in Non-Thermal Plasma Teckniquesfor Pollution Control, Eds. B. M. Penetrante and S. E. Schultheis, Part B, pp, 309-320,1993,

IEEE Transactions on Dielectrics and Electrical Insulation


1991 G. Dinelli and M. Rea, "Pulse Power Electrostatic Technologies for the Control of Flue gas Emission", J. Electrostatics, Vol. 25, pp. 2340,1990. [loa] Y. L. M. Creygton, E. M. Veldhuizen and W. Rntgers, "Electrical and Optivan R. cal Study of Pulsed Positive Corona", in Non-Therinnl Plasma Techniquesfor Pollution Canlrol, Eds. B. M. Penetrante and S. E. Schultheis, Part A, Springer-Verlag,pp. 205230,1993. [loll K. Kawamura, T. Tsukamoto, T. Takashita, S. Katsnki and H. Akiyama, "NO, Removal Using Inductive Pulsed Power Generator", Trans. IEEJ, Vol. 117-A, pp. 956961,1997 (In Japanese) I1021 A. I? Napartovic, Yu S. Akishev, A. A. Deryugin, I. V. Kochelev and N. I. Trushkin, "DC Glow Discharge with Fast Flow for Flue Gas Processing", in Non-Tlieriiial Plasmn Techniquesfor Pollution Control, Eds. B. M. Penetrante and S. E. Schultheis, Part B, Springer-Verlag,pp. 355-370,1993,

Vol. 7 N o . 5, October 2000

683

. Kato, T. Takahashi and K.Shimizu, "Nitric Oxide Decomposition in Air [I031 T Oda, T. by Using Nonthermal Plasma processing with Additives and Catalyst", IEEE Trans. Indnst. Appl., Vol. 34, pp. 268-272,1998.
11041 K. Kato, Y. Kasuga, M. Fujiwara and K. Onda, "Change of Characteristics of Desulfirization and Denitrification by Combustion Flue Gas Composition and Electrode Configuration Under Pulsed Corona Discharge", Electrical Engineering in Japan, Vol. 116, pp, 96108,1996, Translated from Denki Gakkai Ronbunshi, Vol. 115 D, pp. 1046-1053,1995, [lo51 0. Tokunaga, H. Namba and N. Snzuki, "Enhancement of Removal of SO2 and NO, by Powdery Materials in Radiation Treatment of Exhaust Gas", Int. J. Appl. Radiat. Isot., Vol. 36, pp, 807-812,1985.

[lo61 B. Eliasson and U. Kogelschatz, "Nonequilibrinm Volume Plasma Chemical Processing", IEEE Trans. Plasma Sci., Vol. 19, pp, 1063-1077,1991, [I071 H. Yoshida, Z. Marui, M.Aoyama, J. Sngira and A. Miznno, "Removal of Odor Gas Component Utilizing Plasma Chemical Reactions promoted by the Partial Discharge in a Ferroelectric Pellet Layer", IES Japan, Vol. 13, pp, 425430,1989. [lo81 B. M. Penetrante, M. C. Hsiao, J. N. Bardsley, B. T. Merritt, G. E. Vogtlin and l? H. Wallman, "Electron Beam and Pulsed Corona Processing of Volatile Organic Compounds and Nitrogen Oxides", 10th hit. Pulsed Power Conf., Vol. 1, pp. 144150, 1995. [lo91 T. Namihira, S. Tsukamoto, D. Wang, S. Katski, R.Hackam and H. Akiyama, "Axial Distribution of NO, in a Coaxial Reactor During NO, Removal With Pulsed Power", 13th Int. Conf. Gas Discharges and Their Applications, Glasgow, 3-8 September, pp. 1-4,2000, [I101 G. E. Vogtlin and B. M. Penetrante, "Pulsed Corona Discharge for Removal of NOx from Flue Gas", in Non Thermal PInsIna Techniquesfor Pollution Control, Eds. B. M. Penetrante and S. E. Schultheis, Part B, pp. 187-198,1993, [lll] Y. S. Mok, S. W. and 1. S. Nam, "Evaluation of Energy Utilization Efficiencies Ham for SO2 and NO Removal by Pulsed Corona Discharge Process", Plasma Chemistry and Plasma Processing, Vol. 18, pp. 535-550,1998, 11121 R. Korzekwa and L. Rosocha, "A High Temperature Pulsed Corona Plasma Reactor'', 10th IEEE Int. Pulsed Power Conf. pp, 138-143,1995, 11131 N. W. Frank and S. Hirano, "The Electron Beam Flue gas Treatment Process", Radiat. Phvs. Chem., Vol. 35, PD. 416421,1990.
.I

Manuscript was received on 12June 2000.

Vous aimerez peut-être aussi