Investigation of Nanosecond-pulsed Dielectric Barrier Discharge Actuators with Powered. Electrode of Different Exposures

Similar documents
Electric Field Measurements in Atmospheric Pressure Electric Discharges

Modeling and Simulation of Plasma-Assisted Ignition and Combustion

Energy conversion in transient molecular plasmas:

A KINETIC MODEL FOR EXCIMER UV AND VUV RADIATION IN DIELECTRIC BARRIER DISCHARGES*

Plasma Modeling with COMSOL Multiphysics

Burst Frequency effect of DBD Plasma actuator on the control of separated flow over an airfoil

Influence of the Microplasma Actuator Electrode Configuration on the Induced EHD Flow

Formation of white-eye pattern with microdischarge in an air. dielectric barrier discharge system

Modeling and Simulation of Plasma-Assisted Ignition and Combustion

Investigation of fluid flow around a cylinder with EHD actuation on inclined plates behind the cylinder

Numerical Study on Influences of Barrier Arrangements on Dielectric Barrier Discharge Characteristics

EXPERIMENTAL STUDY OF SHOCK WAVE INTERACTING PLANE GAS-PLASMA BOUNDARY

PIC/MCC Simulation of Radio Frequency Hollow Cathode Discharge in Nitrogen

Surface corona discharge along an insulating flat plate in air applied to electrohydrodynamically airflow control : electrical properties

Miniaturized Argon Plasma: Neutral Gas Characteristics in Dielectric Barrier Discharge

On the Introduction of the Irreversibility in a DBD Plasma Based Channel Flow: A Study on Entropy Generation Rate

A Multi-Dimensional Limiter for Hybrid Grid

MAPPING OF ATOMIC NITROGEN IN SINGLE FILAMENTS OF A BARRIER DISCHARGE MEASURED BY TWO PHOTON FLUORESCENCE SPECTROSCOPY (TALIF)

Numerical investigation of nanosecond pulsed plasma actuators for control of shock-wave/boundary-layer separation

Numerical Simulation of Flow Separation Control using Multiple DBD Plasma Actuators

Compressible Flow LES Using OpenFOAM

Multidimensional Numerical Simulation of Glow Discharge by Using the N-BEE-Time Splitting Method

A Magnetohydrodynamic study af a inductive MHD generator

Numerical Simulation of Townsend Discharge, Paschen Breakdown and Dielectric Barrier Discharges Napoleon Leoni, Bhooshan Paradkar

Extremely far from equilibrium: the multiscale dynamics of streamer discharges

Theoretical analysis of ion kinetic energies and DLC film deposition by CH 4 +Ar (He) dielectric barrier discharge plasmas

COMPUTATIONAL STUDY OF SEPARATION CONTROL MECHANISM WITH THE IMAGINARY BODY FORCE ADDED TO THE FLOWS OVER AN AIRFOIL

Effect of Gas Flow Rate and Gas Composition in Ar/CH 4 Inductively Coupled Plasmas

Electron Current Extraction and Interaction of RF mdbd Arrays

Ionization Detectors

Computational modelling of Dielectric Barrier Discharge Plasma Actuators for Aerospace Flow control applications

CALCULATION OF SHOCK STAND-OFF DISTANCE FOR A SPHERE

Multiple microdischarge dynamics in dielectric barrier discharges

Wall treatments and wall functions

FLOW CONTROL USING DBD PLASMA ON BACKWARD-FACING STEP

The Effect of Discharge Characteristics on Dielectric Barrier Discharges According to the Relative Permittivity

Two-dimensional Numerical Simulation of a Planar Radio-frequency Atmospheric Pressure Plasma Source

ARTIFICIAL TURBULIZATION OF THE SUPERSONIC BOUNDARY LAYER BY DIELECTRIC BARRIER DISCHARGE

Chap. 5 Ion Chambers the electroscope

1.3 Molecular Level Presentation

Study of airbreathing electric thruster for nearspace propulsion

An Investigation into the Sensory Application of DBD Plasma Actuators for Pressure Measurement

Comsol Multiphysics 在低溫電漿模擬之應用 ~My Little Journal with Simulation

Influence of water vapour on acetaldehyde removal efficiency by DBD

Modeling of Degradation Mechanism at the Oil-Pressboard Interface due to Surface Discharge

SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS

Multi-fluid Simulation Models for Inductively Coupled Plasma Sources

Formation and Long Term Evolution of an Externally Driven Magnetic Island in Rotating Plasmas )

SHEAR-LAYER MANIPULATION OF BACKWARD-FACING STEP FLOW WITH FORCING: A NUMERICAL STUDY

SCALING OF PLASMA SOURCES FOR O 2 ( 1 ) GENERATION FOR CHEMICAL OXYGEN-IODINE LASERS

Electrohydrodynamic phenomena in atmospheric discharges : application to airflow control by plasma actuators

Adaptive numerical simulation of Streamer. Shailendra Singh

Lecture 6: High Voltage Gas Switches

Matti Laan Gas Discharge Laboratory University of Tartu ESTONIA

Numerical Simulation of Atmospheric-pressure Non-equilibrium Plasmas: Status and Prospects

CHARACTERISATION OF NS-DBD PLASMA ACTUATORS FOR SUPERSONIC FLOW CONTROL

MODELING OF PLASMA ACTUATOR AND ITS EFFECT ON FLOW FIELD AROUND RECTANGULAR CYLINDER

3. Gas Detectors General introduction

EHD flow produced by positive and negative point-to-plate corona discharges

Effect of Applied Electric Field and Pressure on the Electron Avalanche Growth

[N175] Development of Combined CAA-CFD Algorithm for the Efficient Simulation of Aerodynamic Noise Generation and Propagation

Nanosecond-scale Processes in a Plasma Pilot for Ignition and Flame Control

Effect of small amounts of hydrogen added to argon glow discharges: Hybrid Monte Carlo fluid model

Ionization Detectors. Mostly Gaseous Detectors

Nanosecond pulse surface discharges for high-speed flow control

DEPARTMENT OF ELECTRICAL ENGINEERING DIT UNIVERSITY HIGH VOLTAGE ENGINEERING

Simulation of unsteady muzzle flow of a small-caliber gun

arxiv: v1 [physics.plasm-ph] 10 Nov 2014

MULTIPACTOR ON A DIELECTRIC SURFACE WITH LONGITUDINAL RF ELECTRIC FIELD ACTION

Hybrid Resistive-Capacitive and Ion Drift Model for Solid Gas Dielectrics

SECONDARY ELECTRON EMISSION OF CERAMICS USED IN THE CHANNEL OF SPT

Fast Biofluid Transport of High Conductive Liquids Using AC Electrothermal Phenomenon, A Study on Substrate Characteristics

Dense plasma formation on the surface of a ferroelectric cathode

Effect of High Voltage Impulses on Surface Discharge Characteristics of Polyethylene

Modeling of a DBD Reactor for the Treatment of VOC

Direct Modeling for Computational Fluid Dynamics

MULTIGRID CALCULATIONS FOB. CASCADES. Antony Jameson and Feng Liu Princeton University, Princeton, NJ 08544

Inplane Microwave Plasma

Turbulence Modeling I!

Active drag reduction in a turbulent boundary layer based on plasma-actuatorgenerated streamwise vortices

Breakdown processes in metal halide lamps

SW103: Lecture 2. Magnetohydrodynamics and MHD models

Modelling of low-temperature plasmas: kinetic and transport mechanisms. L.L. Alves

CHARACTERISTICS OF ELLIPTIC CO-AXIAL JETS

Department of Aerospace Engineering and Engineering Mechanics, The University of Texas at Austin, Austin, Texas 78712, USA

Numerical simulation of active flow control using plasma actuators is dependent

Numerical Simulation of Microwave Plasma Thruster Flow

Current sheath formation in the plasma focus

Influence of surface charges on the structure of a dielectric barrier discharge in air at atmospheric pressure: experiment and modeling

CONTROL OF UNIFORMITY IN CAPACITIVELY COUPLED PLASMAS CONSIDERING EDGE EFFECTS*

Residual resistance simulation of an air spark gap switch.

THE APPROACH TO THE ANALYSIS OF ELECTRICAL FIELD DISTRIBUTION IN THE SETUP OF PAPER INSULATED ELECTRODES IN OIL

Nanosecond Pulse Ionization Wave Discharges on Liquid Surfaces: Discharge Development and Plasma Chemistry

Table of Content. Mechanical Removing Techniques. Ultrasonic Machining (USM) Sputtering and Focused Ion Beam Milling (FIB)

Prospects for High-Speed Flow Simulations

MAGNETIC NOZZLE PLASMA EXHAUST SIMULATION FOR THE VASIMR ADVANCED PROPULSION CONCEPT

Computational and Experimental Analysis of Mach 5 Air Flow over a Cylinder with a Nanosecond Pulse Discharge

A NUMERICAL STUDY ON THE FLOW AND HEAT TRANSFER FOR THE INSIDE OF A NEW DIVERSION-TYPE LNG HEATING DEVICE

EHD gas flow in electrostatic levitation unit

Zonal hybrid RANS-LES modeling using a Low-Reynolds-Number k ω approach

Transcription:

Plasma Science and Technology, Volume 19, Number 9 http://iopscience.iop.org/article/10.1088/2058-6272/aa6f59/meta Investigation of Nanosecond-pulsed Dielectric Barrier Discharge Actuators with Powered Electrode of Different Exposures Shuangyan XU 1 ( 徐双艳 ), Jinsheng CAI 1 ( 蔡晋生 ), Yongsheng LIAN 2 ( 练永生 ) 1 School of Aeronautics, Northwest Polytechnical University, Xi'an, 710072, China 2 Mechanical Engineering Department, University of Louisville, Louisville, KY, 40292, U.S.A Abstract: Nanosecond-pulsed dielectric barrier discharge actuators with powered electrodes of different exposures were investigated numerically by using a newly proposed plasma kinetic model. The governing equations include the coupled continuity plasma discharge equation, the drift-diffusion equation, electron energy equation, Poisson s equation, and the Navier-Stokes equations. Powered electrodes of three different exposures were simulated to understand the effect of surface exposure on plasma discharge and surrounding flow field. Our study showed that the fully exposed powered electrode resulted in earlier reduced electric field breakdown and more intensive discharge characteristics than partially exposed and rounded-exposed ones. Our study also showed that the reduced electric field and heat release concentrated near the right upper tip of the powered electrode. The fully exposed electrode also led to stronger shock wave, higher heating temperature, and larger heated area. Keywords: Nanosecond-pulsed Dielectric Barrier Discharge, plasma actuators, powered electrode, shock wave, flow field. 1. Introduction Dielectric Barrier Discharges (DBD) have been used in various industrial applications, such as nonequilibrium magneto hydrodynamic devices [1, 2], enhancing ignition and plasma-assisted combustion [3, 4], and flow control actuators [5-7]. Extensive studies have been conducted to understand the fundamental mechanism and to improve DBD plasma actuators efficiency [8-10].These studies have shown that the shape of electrodes can significant affect the performance of DBD plasma actuators [11-13]. For example, serrated plasma actuators can induce stream-wise and spanwise vortices in the shear layer [12, 14, 15], which enhance momentum-transfer across the boundary layer and hence increase DBD actuator efficiency [11, 12]. A serrated 1

exposed electrode can also induce larger body force than a conventional trip configuration [16]. Previous studies have also shown that thrust produced by the DBD plasma actuator strongly depends on the thickness of the exposed electrode[17]: a 50% increase in body force was reported when a thin sheet was used as the exposed electrode [18]. DBD plasma actuators are promising for low speed flow control but their applicability to high speed flow control is limited because of the relatively low flow velocity they induce. It has been demonstrated that Nanosecond-pulsed dielectric barrier discharges (NS-DBD) are able to induce flow reattachment at velocities up to Mach number of 0.85 [6]. Unlike DBD plasma actuators that only transfer momentum to surrounding fluids, NS-DBD plasma actuators can also transfer thermal energy. The rapid heating over a short time period generates compression waves or even shock waves near the powered electrodes. Research has been conducted to improve the efficiency of NS-DBD plasma actuators by changing the dielectric thickness [19-21] or actuator geometry [7, 21, 22]. Moreau et al.[7], for example, used a three-electrode configuration to increase the extension length up to the electrode gap, resulting in a 300% increase in the input energy from a current collected at the third electrode. Others [5, 20, 23] applied positive polarity pulses to investigate the effects of actuator parameters on the characteristics of gas heating generated in NS-DBD. Results from these studies suggested that dielectric thickness has little effect on flow control. In Ref. [24], the authors studied experimentally the Q-V Lissajous Figures in NS-DBD, and the results show that the transported charge and energy per pulse are proportional to the applied voltage amplitude, while the transported charges proportionally decrease as the electrode gap increases. In Ref [25], the authors investigated the effect of grounded electrode s width on electrical characteristics of NS-DBD and the results show that the grounded electrode s width has no obvious influence on the plasma light intensity. The exposure of electrodes can affect the momentum and energy transfer to surrounding fluids but limited studies have been devoted to this subject. In our previous work [26], we proposed a simplified plasma kinetic model to investigate the mechanism of the plasma aerodynamic actuation driven by NS-DBD, and the new model was validated efficiently for complicated actuator geometries and problems in multiple dimensions by comparing the results with experimental and numerical data. In this paper we will investigate the effect of powered electrode exposure on NS-DBD plasma actuators. The rest of the paper is structured as follows: First, 2

we present the physical model and numerical method. Then we simulate NS-DBD actuators with three different exposed cases. We compare the discharge characteristics and responses of the induced flow field among these three actuators. Finally we conclude the significance of the right edge and upper tip of the powered electrode on the performance of the fluid flow. 2. Governing equations and numerical method The governing equations for plasma discharge shown below include the continuity equation for each species, the drift-diffusion equation, and the Poisson equation for electric field [27, 28]. n j t Γ R (1) j i ij Γ j j n j E D j n j (2) qn E (3) 0 r i j j where, n j is the species number density, Γ j the corresponding flux, R ij the local creation loss rate due to chemical reactions, E the electric field, q i the corresponding species charge, 0 the permittivity of vacuum space, r the relative permittivity of air, j the electron mobility, and D j the electron diffusivity. The mobility and diffusion coefficient data for the neutral and ion species can be found in the literature [29-32]. The kinetic mechanism and corresponding reaction rate coefficients are listed in Table 1,whose detail explanation can be found in our previous work [26], in which the kinetic plasma model was simplified into one with 7 species and 9 reactions. The species include electron (e - ), electronically excited states of nitrogen (N 2 (B 3 ), N 2 (C 3 )), ions (N 2+, O 2+ ) and ground state neutrals (N 2, O 2). The reaction mechanism includes electron impact reactions (1-2, 6-7), electron-ion recombination reactions (4, 5), quenching reactions (3, 9) and charge exchange reactions (8). Besides, the rates of reactions 6-7 ( k excit ( i )) can be found in Ref. [26]. 3

Table 1 Reaction mechanism. Number densities in m -3, temperature T in K, electron temperature Te in ev, one-body rates in s -1, and two-body rates in m 3 s -1 No. Reaction Rate 1 e - +N 2=>e - +e - +N 2 + 4.1 10-14 exp(-18.7/t e) 2 e - +O 2=>e - +e - +O 2 + 2.3 10-15 T e 1.03 exp(-12.29/t e) 3 N 2(C 3 )=> N 2(B 3 ) 3 10 7 4 e - +N 2+ =>N 2 2 10-7 (300/T e) 0.5 5 e - +O 2+ =>O 2 2 10-7 (300/T e) 0.5 6 e - +N 2=>e - +N 2(B 3 ) 7 e - +N 2=>e - +N 2(C 3 ) k ( ) excit i k ( ) excit i 8 O 2+N 2+ =>N 2+O 2 + 6.0 10-11 (300/T) 0.5 9 N 2(C 3 )+N 2=>N 2+ N 2(B 3 ) 1.0 10-11 The following electron energy equation (EEE) was adopted to obtain the electron energy, n t Γ eγ E Q (4) ε ε e ε Γ n E D n (5) ε ε ε ε ε here, n ε is the electron energy density, which is the multiplier of the electron number density and the mean energy of the species of emitted electron, Γ ε electron energy flux, ε the electron energy mobility and D ε the electron energy diffusion coefficient, and Q ε is the energy loss of electrons due to elastic and inelastic collisions. The electron energy mobility and energy diffusivity are computed using Einstein s relation for a Maxwellion EEDF ε 53 e and D ε ε T e. NS-DBD plasma aerodynamic actuation transfers not only momentum but energy from plasma discharges to surrounding gas flow. The equations for the whole gas dynamics were based on the compressible Navier- Stokes equations including the continuity equation, momentum equation and energy equation with source terms obtained from the discharge plasma: 4

( u ) 0 t (6) ( u) ( u u pi τ) F t (7) ( Cp T) [ ucp T k T] Q t (8) here, ρ is the density, p is the pressure, τ is the viscous shear stress tensor, T is the temperature, C p describes the amount of heat energy required to produce a unit temperature change in a unit mass, and k is the thermal conductivity. F is the body force generated by nanosecond plasma actuator given by F E, where is the space charge density given by i qn i i. Q p u j E, where 5 j E is the heat source term, j is the current density ( j e( Γ + Γ + Γ N O e ) ), and is the viscous dissipation. Usually, compared to the gas heating 2 2 released to the gas flow from the discharge process, the viscous dissipation of the NS-DBD plasma discharge has a negligible influence on the fluid field in such a short pulse discharge time and it was ignored during the simulation in the present model. Figure 1 shows the procedures to solve the problem. The simulation starts from the plasma discharge model and ends with the induced flow dynamic model. For computational efficiency, the calculations were carried out in two stages. The first stage encompassed the first 100 ns of the pulse discharge. For this stage, the discharge model and the fluid dynamic model were fully coupled. The plasma discharge model (Eq.1-4) calculates the averaged statistical properties of the plasma particles by solving the continuity (Eq.1) and drift-diffusion equations (Eq. 2) using the Scharfetter-Gummel method [33] for flux calculation with an adaptive time step. The Poisson equation (Eq.3) for electric field was solved using the LU triangular matrix decomposition [27]. The space charge density and current density from equations (1, 2, 4) and the electric potential from the equation (3) were taken as power input to the flow dynamic model with Navier-stokes equations (6-8), which were solved using the implicit, second-order upwind scheme [28, 34].The temperature and the velocity of the flow field affects the motion of the particles and the reaction rates in the discharge progress. Since the electric effects and particle motion become negligible after the input pulse dies away, in the second stage of the computations, only the fluid dynamic model was employed.

During the simulation, one set of unstructured mesh with about 10 4 grid cells were adopted to couple the discharge model and fluid dynamic model. In order to capture the discharge features near the right edge and upper tip of the powered electrodes, the minimum value of the grid size was taken as 10-7 mm in the simulation. However, for computational efficiency, the time step for discharge process is set 10-11 s while 10-7 s after discharge. Figure 1. The overall procedure to solve the system of equations For each simulation, the initial number density ( n j ) of electron is about 10 14 m -3, the mole fraction of the neutral species is 1 10-10, and the mole fraction for each ion species is set in a way that the space charge is zero under the standard atmospheric pressure and temperature of 300 K. The initial time step is 10-13 s and the maximum time step is 10-9 s during the simulation. Standard boundary conditions for the charged particles as well as the neutral particles described in Ref. [34] were employed: at the wall, the ion flux was set to zero and the ion temperature was same as the wall temperature, where the ion velocity was set to satisfy the secondary emission with secondary electron emission coefficient 0.01 and electron temperature was set 2 ev. At the far-fields, the normal derivatives of the ion and electron flow to open boundaries are set to zero. The surface charge accumulation was determined by integrating 6

t n j n j, where s ion e s is the surface charge density, j ion is the ion current density, j e is the electron ion density. The dielectric layer was approximated to a uniform electric field E d, which is related to the internal electric field at the surface through the relation n ( 0 red 0E ) s. No-slip boundary conditions with a constant temperature wall were set for the hydrodynamics model. The far-fields boundary conditions of Navier- Stokes equations are set to be open boundaries. 3. Method validation Plasma discharge and flow actuation on a typical asymmetric NS-DBD plasma actuator were simulated under the conditions described in [6] by solving the five-moment equations for particle discharge and Navier- Stokes equations for flow field in our previous work [26]. In this paper, under the same conditions, the NS-DBD plasma actuation were carried out by solving the drift-diffusion equations for particle discharge using the simplified kinetic chemical model. As shown in Figure 2, the powered electrode, with three surfaces (top, left and right) exposed, is 0.0625 mm wide and 5 mm long. The powered and ground electrodes are separated by a fluorocarbon film with a thickness of 0.3 mm. The exposed electrode was powered by a single-pulsed highvoltage with a peak amplitude of 12 kv. The waveform of the applied voltage is shown in Figure 2. The pulse duration is 60 ns, and the pulse rising time is about 15 ns. A large computational domain of 20 mm by 10.1 mm was used to avoid any effect from the outer boundary. Energy released from the discharge leads to a rapid pressure growth in the adjacent gas layer, resulting in a series of cylindrical shaped shock waves. The shock wave is formed at the moment of initial expiation and starts to move toward the surface. The computed wave front positions are compared with experimental results [6] in Figure 3, showing a good agreement, with less than 5% difference. During discharge, the energy input is deposited along the surface above the grounded electrode in the dramatically changing electric field. 7

Figure 2. Diagram of computational domain (left) and waveform of the applied voltage Figure 3. Comparison of the shock wave front location with experimental results [6] 4. Results and Discussion To understand the effect of surface exposure of powered electrodes, we study three configurations of NS- DBD plasma actuators (shown in Figure 4). In all the three configurations, the powered and grounded electrodes are separated by a 0.0375 mm thick dielectric. The grounded electrodes are rectangles of 5 mm by 0.0625 mm. In Case A, only the upper surface of the powered electrode is exposed (partially exposed electrode). The partially exposed electrode has the same dimensions as the grounded electrode. In Case B, three surfaces of the powered electrode are exposed (fully exposed electrode). The exposed electrode has the same geometry as the grounded electrode. In Case C, three surfaces of the powered electrode are exposed, but the exposed powered electrode 8

has a round semi-circular corner with a 0.00375mm diameter (rounded-exposed electrode). In all cases, the powered electrodes were driven by a single positive nanosecond-pulsed high-voltage (shown in Figure 5). The parameters are fixed in three cases during the simulation. The single pulse has an amplitude of 6 kv, a rising time of 40 ns and pulse duration of 80 ns. The half band width of the pulse was about 40 ns. Figure 4. Diagram of the three NS-DBD plasma actuators. Case A has a rectangular powered electrode with only exposed upper surface; Case B has a rectangular powered electrode with three exposed surfaces; Case C has round cornered powered electrode with three exposed surfaces. Figure 5. Waveform of the pulsed voltage used in the simulation. 4.1 Discharge characterization 9

Figure 6 compares the spatial distributions of the reduced electric field ( E / N n, N n is the density of the neutral gas) among the three cases at 40 and 80 ns, which correspond to the peak and end of the input voltage pulse (see Figure 5), respectively. In Figure 6, we used the same range scale in three cases, and all the values greater than 450 Td are represented by the same color as that of 450 Td so that the discharge domain can be resolved clearly. It is found that the streamers start from the region near the right upper tip of the powered electrodes. As the input signal increases, the streamers run right along the dielectric surface above the grounded electrodes and the discharge domain expands. Under the influence of the reduced electric field, the distributions of the species vibration temperature are similar to those of the reduced electric field for three cases. Despite these common features, in Figure 6, at the instant of peak input voltage (40 ns), the electric breakdown area in Case B is 0.1 mm longer and 0.05 mm higher than that created in Case A with the partially exposed electrode. The induced area in Case B is the largest with the strongest reduced electric field, while that in Case C is the second largest with the second strongest reduced electric field and that in Case A is the smallest with the weakest electric field. At 80 ns, when the input signal decays to zero, the electric field decreases to zero in Case A, but is visible around the right edges of the powered electrodes in Case B and C. The comparison demonstrates that the reduced electric field in Case B is stronger than Case A and C, and the exposure effects of the powered electrode are important to the transfer of momentum and energy into the surrounding air. 10

Figure 6. Comparison of the reduced electric field at 40 ns (Left) and 80 ns (Right). First row: Case A; Second row: Case B; Third row: Case C. We further compare the contours of the electron density and power density at 40 ns in Figure 7. In Case A with partially exposed electrode, electrons are concentrated at the upper tip of the anode, where the values of the electron density and power density are the highest. In cases B and C, electrons are concentrated near the right upper tip of the powered electrode throughout the discharge process primarily due to the influence of the electric field. As a result, the electron density forms a core discharge domain near the right upper tip of the powered electrode with higher density values. The streamers run right along the surface above the grounded electrode, accompanied by a reduction of the electron density. Table 2 presents the comparison of the sizes of the circumscribed rectangle of the streamers of the three cases at 40 ns, respectively. For electron density, the area of the circumscribed rectangle in Case B is approximately 4 times bigger than that in Case A, and 2 times bigger than that in Case C. For power density distributions, the area of the circumscribed rectangle in Case B is 7 times bigger than that in Case A, and 3 times bigger than that in Case C. This difference is primarily due to the higher reduced electric field profile for exposed right edge and tip of the powered electrode. 11

Figure 7 Distribution of electron density (Left) and power density (Right) (a) Case A (b) Case B (c) Case C Table 2 Comparison of the circumscribed rectangles of the discharge streamers at 40 ns Length (mm) Height (mm) Variables Case A Case B Case C The streamer where the Electron 0.1 0.05 0.2 0.12 0.1 0.07 density is higher than 1.4 10 18 1/m 3 The steamer where the Power density is higher than 3 10 12 1/m 3 0.05 0.03 0.2 0.06 0.1 0.04 Figure 8. Temporal distributions of (a) electric body force and (b) power density for three exposed cases 12

Figure 8 compares temporal profiles of the electric body force and power density for three exposed cases. Due to particle collisions happened around the exposed electrodes (Case B and C) in the high electric field, higher space charge density and current density are generated. The temporal profiles of body force and powered density for Case B and C are similar, both are higher than those for Case A. From the numerical results, it can be concluded that the complete exposure electrodes can increase the induced energy-transfer performance and momentum-transfer compared with those induced by partial exposure electrode. 4.2 Characterization of the induced flow The 3 cases were simulated by coupling the plasma kinetic model with the Navier-Stokes equations to study the effect of actuators on surrounding neutral gas. In the flow control by NS-DBD plasma actuator, energy release leads to a pressure increase in the gas layer near the surface of the dielectric. Figure 9 shows the pressure waves along a horizontal line of y = 0.6 mm and a vertical line of x = 9 mm (immediately on the right edge of the powered electrode) within 6 µs. Right after the electric field breakdown, a strong pressure wave was generated at 0.1 µs. The wave strength decays as it propagates. Case B has the strongest pressure wave at 0.1 µs while Case A has the weakest. Note that, at the same instant, the waveform in Case B is fatter than that in the other two Cases, which might be due to the higher reduced electric field in Case B. For all cases, the pressure wave expands with a velocity of 0.3 mm/µs along both the x and y directions, giving rise to a shock of M=1.23 near the surface. 13

Figure 9. The evolution of pressure wave along a horizontal line y = 0.6 mm (left) and a vertical line x = 9 mm (right). (a) Case A; (b) Case B; and (c) Case C. In order to investigate the effects of the right upper tip and edge of the powered electrode on the induced flow, Figure 10 presents the evolution of the pressure waves at 1, 3 and 5 µs after the discharge induced by Case A and Case B. In both cases, it is obviously that fast discharge thermalization merges into a single cylindrical compression wave centered on the right upper electrode tip in these two cases. The compression waves decay into weak pressure waves as they propagate with time. However, the pressure generated in Case B propagates slightly faster than those induced in Case A, which is due to the more intensive wave in the Case B. The comparison of the results demonstrates that the exposure of the right upper tip and edge of the powered electrode significantly affect the propagation of the pressure waves. 14

Figure 10. Comparison of pressure waves between Case A (Left) and Case B (right) at 1µs, 3µs and 5µs. Figure 11 shows the temperature distribution at 1 and 5 µs, which is generated by fast gas heating from the discharge. In all cases, the gas heating almost occurs near the right tip and edge of the powered electrode, where electrons concentrate under the high electric field with higher temperature values. During the short pulse discharge time, the energy deposited and released into the surrounding gas. Despite these common features, gas temperatures in Case B and C are larger in heated region and higher in temperature values than those in Case A using the same temperature scale. It indicates that more effective heating was induced with right edge and tip exposed. Furthermore, the highest temperature is observed near the tip of the powered electrodes in Case B. Comparing the temperature distribution at 5 µs between Case B and Case C, the gas heating area is larger in the Case B. The light difference verified the positive affluence of right upper tip on performance of the gas heating. 15

Figure 11. Temperature distributions at 1 µs and 5 µs (a) Case A (b) Case B (c) Case C Since temperatures exhibit similar during the effective actuation period, respectively, profiles of temperature along horizontal line y = 0.503 mm and vertical line x = 9.02 mm at 3 µs are plotted in Figure 12, and temperature at other moment are similar. As shown in Figure 12(a), along the horizontal line y = 0.503 mm, temperature peaked near the right edge (about x = 9 mm) of powered electrodes but immediately dropped as moving away from the tip. It can be seen that the temperature undergoes a drastic rise within the discharge time, which is due to the deposition of the electrons near the right upper tip of the powered electrode. At 3 µs, temperature reaches to about 309 K, 430 K and 405 K for Case A, Case B, and Case C, respectively. In Figure 12 (b), along the vertical line x = 9.02 mm, the heating region ranges from y = 0.5 mm to around y = 0.56 mm with peak value about 309K for Case A. For Case B, temperature heating distributes from y = 0.5 mm to around y = 0.6 mm with a temperature increase of 130 K. For Case C, temperature heating occurs from y = 0.5 mm to around y = 0.58 mm, with temperature increase 115 K. These analytic data demonstrates that the temperature goes down to a value close to ambient temperature after discharge. It is obvious that higher gas heating and larger involved area was induced near the right edge and right upper tip of the powered electrode. Furthermore, the peak temperatures occur immediately near the right edge of the powered electrodes, and the temperature increases most rapidly in Case B, while that in Case C the second rapidly, and that in Case A is the slowest. The 16

comparison results demonstrate that the gas heating temperature significantly depends on the exposure of the powered electrode. Figure 12. Temperature profiles at 3 µs (a) along a horizontal line y = 0.503 mm (b) along a vertical line x = 9.02 mm 5. Conclusions Effects of the powered electrode exposure on air flow induced by three single NS-DBD plasma actuators were investigated in quiescent air at atmospheric pressure. Three exposed cases were studied using a reduced kinetic chemistry model. Discharge characteristics and induced flow features were compared. Our study showed the fully exposed case had earlier electric field breakdown, greater power energy input concentrated at the right edge and upper tip of the powered electrode, and had a longer duration and higher amplitudes than the partially exposed or rounded-exposed cases. Furthermore, the fully exposed electrode had higher heat release and stronger shock wave. Acknowledgments Shuangyan Xu and Jinsheng Cai were supported by the National Natural Science Foundation of China (11472221), and was also funded by 111 project of China. The authors are also grateful to Dr. Y.F. Zhu for his hearty help with this work. 17

REFERENCES 1. Murray, R.C., et al., 2006, Aiaa Journal, 44(1): p. 119-127. 2. Adamovich, I.V., et al., 2008, Journal of Propulsion and Power, 24(6): p. 1198-1215. 3. Pancheshnyi, S.V., et al., 2006, Ieee Transactions on Plasma Science, 34(6): p. 2478-2487. 4. Sun, W.T., et al., 2012, Combustion and Flame, 159(1): p. 221-229. 5. Little, J., et al., 2012, Aiaa Journal, 50(2): p. 350-365. 6. Roupassov, D.V., et al., 2009, Aiaa Journal, 47(1): p. 168-185. 7. Bayoda, K.D., N. Benard, and E. Moreau, 2015, Journal of Applied Physics, 118(6). 8. Benard, N., et al., 2016, Experiments in Fluids, 57(2). 9. Sato, M., et al., 2015, Physics of Fluids, 27(11). 10. Sujar-Garrido, P., et al., 2015, Experiments in Fluids, 56(4). 11. Roy, S. and C.C. Wang, 2009, Journal of Physics D-Applied Physics, 42(3). 12. Liu, Z.F., L.Z. Wang, and S. Fu, 2011, Science China-Physics Mechanics & Astronomy, 54(11): p. 2033-2039. 13. Hoskinson, A.R. and N. Hershkowitz, 2010, Journal of Physics D-Applied Physics, 43(6). 14. Durscher, R.J. and S. Roy, 2012, Journal of Physics D-Applied Physics, 45(3). 15. Joussot, R., et al., 2013, Journal of Physics D-Applied Physics, 46(12). 16. Thomas, F.O., et al., 2009, Aiaa Journal, 47(9): p. 2169-2178. 17. Enloe, C.L., et al., 2004, Aiaa Journal, 42(3): p. 595-604. 18. Abe, T., et al., 2008, Aiaa Journal, 46(9): p. 2248-2256. 19. Correale, G., R. Winkel, and M. Kotsonis, 2015, Journal of Applied Physics, 118(8). 20. Dawson, R. and J. Little, 2013, Journal of Applied Physics, 113(10). 21. Ndong, A.C.A., et al., 2013, Journal of Electrostatics, 71(3): p. 246-253. 22. Takashima, K., et al., 2011, Plasma Sources Science & Technology, 20(5). 23. Dawson, R.A. and J. Little, 2014, Journal of Applied Physics, 115(4). 24. Jiang, H., et al., 2013, Ieee Transactions on Dielectrics and Electrical Insulation, 20(4): p. 1101-1111. 25. Hui Jiang, T.S., Cheng Zhang, Ping Yan, Zheng Niu, Yixiao Zhou, in 2013 Annual Report Conference on ELectrical Insulation and Dielectric Phenomena. 2013.DOI:10.1109/CEIDP.2013.6747419 26. S.Y.Xu, J.S.C., and J.Li, 2016, Physics of Plasmas, 23: p. 103510. 27. Bak, M.S. and M.A. Cappelli, 2013, Journal of Applied Physics, 113(11). 28. Zhu, Y.F., et al., 2013, Journal of Physics D-Applied Physics, 46(35). 29. Deconinck, T., S. Mahadevan, and L.L. Raja, 2007, Ieee Transactions on Plasma Science, 35(5): p. 1301-1311. 30. Mahadevan, S. and L.L. Raja, 2010, Journal of Applied Physics, 107(9). 31. Nelson, D., et al., 2003, Journal of Applied Physics, 94(1): p. 96-103. 32. Viehland, L.A. and E.A. Mason, 1995, Atomic Data and Nuclear Data Tables, 60(1): p. 37-95. 33. Chawla, B.R. and H.K. Gummel, 1970, Ieee Transactions on Electron Devices, Ed17(10): p. 915-&. 34. Poggie, J., et al., 2013, Plasma Sources Science & Technology, 22(1). 18