Open access peer-reviewed chapter

Active Flow Control by Using Plasma Actuators

Written By

Gabriele Neretti

Submitted: 30 September 2015 Reviewed: 25 February 2016 Published: 08 September 2016

DOI: 10.5772/62720

From the Edited Volume

Recent Progress in Some Aircraft Technologies

Edited by Ramesh K. Agarwal

Chapter metrics overview

3,139 Chapter Downloads

View Full Metrics

Abstract

Active flow control has recently received an increasing attention since it allows to directly manipulate the flow-field around a surface only when it is effectively requested. Aerodynamic plasma actuators supplied by a dielectric barrier discharge (DBD) can be used for this purpose. Usually, sinusoidal voltages in the range 5–50 kV peak and frequencies between 1 and 100 kHz are utilized to ignite this plasma typology. The surface discharge produced by these devices is able to tangentially accelerate the flow field by means of the electrohydrodynamic (EHD) interaction. DBDs generate non-thermal plasmas characterized by low input energies and limited temperature increments. Plasma actuators can be easily designed by following the shape of the aerodynamic body and can be used over heat-sensitive surfaces. These aerodynamic devices have demonstrated to produce boundary layer modifications with induced speeds up to 10 m/s. Their use over airfoils, flaps, and blades have shown the possibility to delay the transition between laminar to turbulent regime, to prevent flow separation enhancing lift and reducing drag. Moreover, the adoption of these actuators over landing gears and trailing edges may induce a noise reduction effect. Dielectric materials, electrodes configuration, and supplying waveforms are most relevant parameters to be considered to enhance actuator performance. On a parallel plane, on/off actuation strategy is a key point in the use of these devices when utilized over aerodynamic surfaces impinged within an external flow.

Keywords

  • Active flow control
  • Dielectric barrier discharge
  • Electrohydrodynamic interaction
  • Plasma physics
  • Noise mitigation

1. Introduction

The interest in flow control in the fluid dynamic domain has showed a quick growth in last 15 years [1] due to rapid development of sensors and actuators technologies. Even though passive control techniques are still attractive since they do not require an energy input, active control strategies have recently received more attention since they can be used in a selective way and can be operated only when it is effectively requested. Among different active techniques, plasma aerodynamic actuators are attractive because they present high dynamic responses due to the absence of moving parts, are characterized by low weight, are easy to build, are backward compatible with existing aerodynamic surfaces, and generate negligible aerodynamics interferences when they are switched off. When actuated, they can significantly modify the status of the boundary layer developing on the body surfaces. For this reason, they have been extensively studied for aeronautical applications to prevent flow separation enhancing lift and reducing drag [2, 3]. More recently, they have been used also to control friction drag by delaying transition [4, 5] or by oscillating the flow in spanwise direction [6], and to control global instabilities of the flow [2, 3, 7]. Due to these characteristics, the potential of plasma actuators has been extended to many other applications like for instance tip clearance flow control of turbines [8, 9], and wind turbine blades and holder [10, 11].

This brief review about DBD aerodynamic actuators is all but exhaustive. The main goals of this work are to give to the reader some basic knowledge about the physics of plasma actuators and to show most important applications of these devices into the active flow control domain.

Advertisement

2. DBD aerodynamic actuators: basic principles

Plasma aerodynamic actuators are based on the electrohydrodynamic (EHD) interaction generated by the so-called ionic wind. High electric fields can locally ionize the air. The produced heavy-charged species are accelerated by the applied electric field and, by means of collisions, they can yield momentum to the surrounding neutral gas. The force f per unit volume within the discharge can be yield by the following expression:

f=nineE¯E1

where E is the electric filed and ni and ne are ion and electron number density respectively.

First studies were conducted by using direct current (DC) corona discharges [2]. Main disadvantages of these actuators are low flexibility in electrodes configuration and in the supply system, and the transition into spark regime producing irreversible damage of the actuator itself.

Nowadays, the largest part of plasma aerodynamic actuators is based on the surface dielectric barrier discharge (SDBD). This typology of discharge allows to use several electrodes geometry, different supply voltage waveforms, and it prevents the transition into the arc regime due to the presence of the dielectric material. DBD actuators generate non-thermal plasmas [12], limiting power consumption and allowing their use over heat sensitive surfaces.

A classical DBD aerodynamic plasma actuator is constituted by a dissymmetric electrode pair separated by a dielectric slab (Figure 1). When high voltages (typically 10–50 kV at 1–100 kHz) are applied to the electrodes, a surface discharge is produced (Figure 2). Discharge appears to be macroscopically homogeneous to the unaided eye, but it is constituted by a sequence of micro-discharges [13] lasting typically for tens of nanoseconds with a repetition rate of several hundreds of megahertz. Fast ignition and quenching of the discharge can be inferred by voltage-current time behavior reported in Figure 3.

Figure 1.

Schematic of a SDBD actuator for flow control.

Figure 2.

Top view image of SDBD discharge [14].

Figure 3.

Voltage-current time behavior of a SDBD [15].

The presence of the plasma and the particular electrode configuration induce a jet tangential to the actuator wall, similar to a classic blowing technique [13]. These jets can modify the aerodynamic boundary layer, increasing flow momentum, at least in the near-wall region above the surface. A large number of experimental [1425] and numerical [2647] works have been done in last decade to understand basic physical phenomena involved in the EHD interaction.

A typical velocity profile induced by the tangential wall jet is shown in Figure 4. Measurements have been carried out by a glass Pitot tube positioned 2 mm downstream with respect plasma extension and moved parallel to the actuator surface. Maximum velocity of about 5–6 m/s are usually reached.

Figure 4.

Typical pitot velocity profile of a DBD plasma actuator (red line). Blue lines represent standard deviation [48].

Fast dynamics of these actuators is underlined in Figure 5 where Schlieren images of the induced jet developing during ignition of the discharge are shown. Steady state (d) is typically reached after 400 ms after discharge ignition.

Figure 5.

Schlieren images of induced wall jet during ignition phase of the discharge (a, b, and c) and steady state operation (d) [48].

As already introduced [2647], EHD interaction produced by DBD plasma actuators has been numerically investigated a well. Main issues are related to the simulation of the discharge and its interaction with surrounding gas. A ‘realistic’ simulation of a DBD streamer should take into account gas ionization, plasma chemistry, thermal fluxes, diffusion of neutral and charged species, and charge deposition over dielectric surfaces. Time steps of about tens of picoseconds and mesh size in the order of few micrometers are usually required to reproduce the formation of a plasma filament lasting for tens of nanoseconds and featured with characteristic length of about few millimeters. Several coefficients related to diffusion, secondary electrons emission, reaction rates, and electrons attachment are often not easy to estimate or to recover from existing literature. The numerical model leads to the temporal/spatial evolution of the body force produced by the discharge. The second step is the coupling between the discharge and the surrounding gas. This is another numerical challenge because characteristic times and lengths of the discharge are several orders of magnitude smaller with respect those related to the fluid-dynamic domain. Several authors [2636] followed this computational strategy obtaining interesting results in good agreement with experimental fluid-dynamic effects. In Figure 6a, integrated horizontal body force as a function of the applied voltage is shown [36]. The two half periods produce different force dynamics and magnitude. In Figure 6b, the time-averaged horizontal force distribution above the actuator surface is depicted [33]. Highest values have been obtained close to plasma end.

Figure 6.

Time evolution of the EHD body force (a) [36], and spatial distribution of the average horizontal force (b) [33].

A simpler approach is to estimate the electron number density [3742], on the basis of experimental measurements or theoretical calculations, and to evaluate EHD body force using Equation 1. This force is subsequently utilized as input parameter in the Navier-Stokes equation solver. A spatial/temporal average value of the body force can also be obtained by means of experimental results [4347]. This method allows to use standard fluid-dynamic solvers avoiding difficulties arising from the couple between plasma and surrounding gas.

Both numerical and experimental works demonstrated the capability of plasma actuators to produce a thrust able to move surrounding gas in a desired direction. In order to increase both induced speed and region in which the actuator manifests its influence, a series of DBD actuators can be manufactured (Figure 7). Each actuator of this multi-electrode arrangement contributes to increase the induced speed.

Figure 7.

Horizontal velocity profile (y = 0.5 mm) above four DBD actuators in series [13].

On a parallel plane, voltage waveform shape (arbitrary or nanopulsed signals) can enhance thrust of the induced jet, increasing effectiveness of actuators [15, 4956]. The use of arbitrary waveforms can enhance plasma ignition phases in which induced thrust is higher. In this way, induced speed (Figure 8) and efficiency (Figure 9) can be both increased [15]. Nanopulsed discharges are able to generate energetic plasmas lasting for few nanoseconds. During the ignition of the discharge, a high amount of energy is deposited within the gas inducing local shock waves able to significantly modify the boundary layer of an incoming flow [56].

Figure 8.

Pitot velocity profiles obtained close to plasma extension for different supplying voltage waveforms [15].

Figure 9.

Electric into kinetic conversion efficiency for a DBD actuator driven with different voltage waveforms [15].

Another possibility to increase actuator performance is the adoption of a third exposed electrode supplied with high-voltage DC fields. In this way, a sliding DBD is generated. Depending on the sign of the DC field, it is possible to modify the morphology of the induced ionic wind [57, 58].

DBD actuators can be arranged to produce normal or vectorized jets, too. These devices can be used to mimic classical synthetic jets and are usually called plasma synthetic jet actuators (PSJA). Typical geometries are annular and linear one [48, 5962]. In both cases, tangential wall jets collide merging in a unique-induced jet able to generate perturbations far away from the actuator surface (Figures 10 and 11). When the linear configuration is adopted, by supplying exposed electrodes with different voltages, it is possible to produce an induced jet with an arbitrary inclination. Three-dimensional flows can be also produced by using a classical DBD actuator, but with an exposed electrode characterized with serpentine or serrated geometries, instead of the usual linear one [63, 64]. This particular electrode geometry generates small tangential jets that collide and propagate downstream pushed by the EHD interaction body force.

Figure 10.

Induced normal jet of linear (left-hand side) and annular (right-hand side) PSJA.

Figure 11.

Schlieren images of PSJAs of linear (a) and annular (b) geometries. Electrode distance (a) and diameter (b) is 30 mm.

Advertisement

3. DBD aerodynamic actuators: flow manipulation

In the last decade, DBD aerodynamic actuators have been extensively studied in the active flow control domain. Position and actuation strategies of these devices are key points in their effectiveness in flow control over aerodynamic surfaces. Many studies have been accomplished over airfoils [6572], diffusers [73, 74], and wind [10, 11 and 75] and turbine [8, 9] blades, in order to enhance fluid-dynamic performance. Moreover, the adoption of these devices over landing gears and trailing edge surfaces have demonstrated the possibility to obtain a noise reduction effect [7678].

DBD actuators can be usually positioned over aerodynamic surfaces in spanwise and streamwise directions. In the former, the induced body force is in the same direction as the incoming flow. In the latter, induced thrust is perpendicular to the free stream direction. In this case, the composition of these two flows produces vorticities propagating in the downstream direction.

In Figure 12, a NACA 0015 airfoil equipped with four spanwise DBD plasma actuators is shown. The four plasma regions are clearly visible in the figure as bluish strips. In this configuration, actuators produce a tangential wall jet directed downstream. The effect of this plasma device in the recovery of stall condition is depicted in Figure 13, where smoke visualization is reported. Experiments show how the most effective actuator is the one positioned on the airfoil leading edge, just before the region where separation occurs [66, 67, 79, 80].

Figure 12.

Airfoil equipped with four spanwise plasma actuators [65].

Figure 13.

Flow structure around NACA0015 airfoil at Re = 15000 without (left-hand side) and with (right-hand side) plasma actuation [65].

When DBD actuators are streamwise mounted, induced and incoming flows combine together producing vorticities propagating in the downstream direction (Figure 14). Such a device can be used as a plasma vortex generator (PVG)

Figure 14.

Vortex formation mechanism induced by a streamwise plasma actuator [8].

Changes in drag coefficient (CD) and lift coefficient (CL) by using a PVG are reported in Figure 15. At low Reynolds numbers, actuator effectiveness is very pronounced, leading to noticeable CD reduction with a parallel increment in the stall angle.

Figure 15.

Drag (a) and lift coefficient enhancement (b) with counter-rotating PVGs on a NACA 4418 airfoil [8].

Figure 16.

Pressure coefficient contours with streamlines over an airfoil with 15° angle of attack: plasma actuator off (a) and on (b) [47].

Figure 17.

Lift recovery in percentage for different actuators location with a free stream velocity of 11 m/s: steady operation (a) and duty cycle operation (b) [66].

Effectiveness in flow control by means plasma actuators has been demonstrated by numerical works, too [47, 8187]. In Figure 16, pressure coefficient contours with streamlines obtained over an airfoil with a 15° angle of attack are displayed. In Figure 16a, plasma actuator is switched off, and separation occurs. When the plasma device is activated, reattachment of the flow is achieved (Figure 16b).

Plasma actuators have demonstrated to strongly improve their ability in flow manipulation when operated with a duty cycle strategy [47, 66, 71, 72, 88]. With this approach, the DBD device is turned on and off with intermittence by following a particular duty cycle frequency. This frequency is usually chosen in the range 5–100 Hz and it is strictly related to natural vorticities developing over the aerodynamic surface. The percentage ratio between the period in which discharge is fed and the whole duty cycle period is called duty cycle percentage. If it is fixed to 50%, it means that discharge is ignited half of the time. On a parallel plane, this actuation strategy leads to lower power consumption.

Figure 17 shows lift recovery in percentage of a stalled NACA0015 airfoil equipped with spanwise vectorized actuators located in different airfoil positions [66]. Jet 5 is generated by a DBD actuator located in the leading edge. When actuators are continuously operated, lift increments are limited to about 15%. When operated with a duty cycle strategy, lift increments are close to 50%.

Plasma actuators can also be used to reduce noise induced by aerodynamic surfaces, especially by landing gears and trailing edges [7678]. Studies on bluff bodies have demonstrated the ability of DBD actuators to reduce downstream turbulence, leading to the suppression of particular tones or to an overall noise mitigation up to 4 dB.

Advertisement

4. Conclusion

Studies carried out in the last decades have demonstrated the possibility to use DBD aerodynamic actuators for active flow control purposes. In this last section, main advantages and disadvantages in the use of these plasma devices will be summarized.

Pros

  • Possibility to locate these actuators in different positions on a surface and over existing aerodynamic bodies.

  • Negligible aerodynamic interferences when they are not active.

  • Fast actuation times.

  • Low power consumption.

  • Possibility to tune the on/off actuation strategy depending on the particular fluid dynamic condition.

Cons

  • Induced velocity below 10 m/s.

  • Low electric into kinetic conversion efficiency.

  • Use of potentially hazardous high voltages and electromagnetic noise generation.

  • Effectiveness in flow control for Reynolds numbers typically below few hundred thousand.

Disadvantages already introduced will be overcome only when a more detailed knowledge of basic interactions between discharge a neutral gas will be achieved. When higher induced speeds will be available, new and more effective flow control strategies will be developed.

References

  1. 1. M. Gad-El-Hak. Flow Control. Cambrige: Cambrige University Press; 2000.
  2. 2. E. Moreau. Airflow control by non-thermal plasma actuator. J. Phys. D: Appl Phys. 2007;40:605. DOI: 10.1088/0022-3727/40/3/S01
  3. 3. J.J. Wang, K. Choi, L. Feng, T. N. Jukes, R. D. Whalley. Recent developments in DBD plasma flow control. Progress in Aerospace Sciences. 2013;62:52–78. DOI:10.1016/j.paerosci.2013.05.003
  4. 4. S, Grundmann, C. Tropea. Experimental transition delay using glow-discharge plasma actuators. Exp. Fluids. 2007;42:653–657.
  5. 5. S. Grundmann, C. Tropea. Active cancellation of artificially introduced Tollmien-Schlichting waves using plasma actuators. Exp. Fluids. 2008;44:795–806.
  6. 6. S.P. Wilkinson. Investigation of an oscillating surface plasma for turbulent drag reduction. In: Proceedings of the 41st aerospace sciences meeting and exhibit; Reno, USA. AIAA; 2003. AIAA 2003–1023
  7. 7. G. Touchard. Plasma actuators for aeronautical applications-state of art review. International Journal of Plasma Environmental Science and Technology. 2008;2(1):1–25.
  8. 8. Choi, K.–S., Jukes, T.N., Whalley, R.D., Feng, L.H., Wang, J.J., Matsunuma, T. and Segawa, T.. Plasma Virtual Actuators for Flow Control. Journal of Flow Control, Measurement & Visualization. 2005;3:22–34. DOI:10.4236/jfcmv.2015.31003.
  9. 9. Thomas C. Corke, Flint O. Thomas, and Junhui Huang. Documentation and Control of Flow Separation on a Low Pressure Turbine Linear Cascade of Pak-B Blades Using Plasma Actuators. In: NASA/CR-2007–214677, editor. University of Notre Dame, Aerospace and Mechanical Engineering Department, . 2007.
  10. 10. Robert C. Nelson Thomas C. Corke, Hesham Othman, Mehul P. Patel, Srikanth Vasudevan, and Terry Ng. A Smart Wind Turbine Blade Using Distributed Plasma Actuators for Improved Performance. In: 46th Aerospace Sciences Meeting; January 7–10; Reno, Nevada (USA). 2008. DOI: AIAA-2008–1312.
  11. 11. D. Greenblatt, A. Ben-Harav, H. Mueller-Vahl. Dynamic Stall Control on a Vertical-Axis Wind Turbine Using Plasma Actuators. AIAA Journal. 2014;52:456–462.
  12. 12. U Kogelschatz. Dielectric-barrier Discharges: Their History, Discharge Physics, and Industrial Applications. Plasma Chemistry and Plasma Processing. 2003; 23(1):1–46.
  13. 13. N. Bénard and E. Moreau. Electrical and mechanical characteristics of surface AC dielectric barrier discharge plasma actuators applied to airflow control. J. Experiments in Fluids. 2014;55:1846:1–43. DOI: 10.1007/s00348–014-1846-x.
  14. 14. A. Cristofolini, C.A. Borghi, and G. Neretti. Charge distribution on the surface of a dielectric barrier discharge actuator for the fluid-dynamic control. J. Appl. Physics. 2013;113(143307)DOI: 10.1063/1.4799159.
  15. 15. C.A. Borghi, A. Cristofolini, G. Grandi, G. Neretti, and P. Seri. A plasma aerodynamic actuator supplied by a multilevel generator operating with different voltage waveforms. Plasma Sources Science and Technologies. 2015; 24(4). DOI: 10.1088/0963–0252/24/4/045018.
  16. 16. C.A. Borghi, M.R. Carraro, A. Cristofolini, G. Neretti. Electrohydrodynamic interaction induced by a dielectric barrier discharge. J. Appl. Physics. 2008;103DOI: 10.1063/1.2888354
  17. 17. A. Cristofolini, G. Neretti, F. Roveda, and C.A. Borghi. Schlieren imaging in a dielectric barrier discharge actuator for airflow control. J. Appl. Physics. 2012;111(033302)DOI: 10.1063/1.3682488J.
  18. 18. Pons, E. Moreau, G. Touchard. Electrical and aerodynamic characteristics of atmospheric pressure barrier discharges in ambient air. In: Proc. ISNTPT; 2004. p. 370–410.
  19. 19. A. Cristofolini, G. Neretti, C.A. Borghi. Effect of the charge surface distribution on the flow field induced by a dielectric barrier discharge actuator. J. Appl. Physics. 2013;114(073303)DOI: 10.1063/1.4817378
  20. 20. G. Neretti, A. Cristofolini, C.A. Borghi, A. Gurioli, and R. Pertile. Experimental results in DBD plasma actuators for air flow control. IEEE Transactions on Plasma Science. 2012;40:1678–1687. DOI: 10.1109/TPS.2012.2191801
  21. 21. F.O. Thomas, T.C. Corke, M. Iqbal, A. Kozlov, and D. Schatzman. Optimization of Dielectric Barrier Discharge Plasma Actuators for Active Aerodynamic Flow Control. AIAA Journal. 2009;47:2169–2178. DOI: 10.2514/1.41588
  22. 22. T.C. Corke, C.L. Enloe, and S.P Wilkinson. Dielectric barrier discharge plasma actuators for flow control. Annual Review of Fluid Mechanics.2010;42: 505 - 529.
  23. 23. G.I. Font, C.L. Enloe, and T.E. McLaughlin. Plasma volumetric effects on the force production of a plasma actuator. AIAA Journal. 2010;48(9):1869 - 1874.
  24. 24. M. Forte, J. Jolibois, E. Moreau, G. Touchard, and M. Cazalens. Optimization of a dielectric barrier discharge actuator by stationary and non-stationary measurements of the induced flow velocity: application to airflow control. Experiments in Fluids.43(6):917 - 928, 2007.
  25. 25. M. Kotsonis. Diagnostics for characterisation of plasma actuators. Measurement Science and Technology. 2015; 26(9): 092001. DOI: 10.1088/0957–0233/26/9/092001.
  26. 26. J.P. Boeuf, Y. Lagmich, and L.C. Pitchford. Contribution of positive and negative ions to the elecrtrohydrodynamic force in a dielectric barrier discharge operating in air. J. Appl. Phys. 2009;106 (023115 )
  27. 27. Hiroyuki Nishida, and Takashi Abe. Validation Study of Numerical Simulation of Discharge Plasma on DBD Plasma Actuator. In: 42nd AIAA Plasmadynamics and Lasers Conference; 27 – 30 June; Honolulu, Hawaii. 2011. AIAA 2011–3913.
  28. 28. Meenakshi Mamunur, Uwe Kortshagen, Douglas Ernie, and Terrence Simon. Plasma actuator simulation: Force contours and dielectric charging characteristics. In: 48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition; 4 – 7 January; Orlando, Florida. 2010. AIAA 2010–1221
  29. 29. Likhanskii, A., Shneider, M., Macheret, S., Miles, R.. Modeling of dielectric barrier discharge plasma actuators driven by repetitive nanosecond pulses. Physics of Plasmas . 2007;14(073501 )
  30. 30. Likhanskii, A. Simulation Tool for Dielectric Barrier Discharge Plasma Actuators at Atmospheric and Sub-Atmospheric Pressures. NASA/CR-2012–217260. 2012.
  31. 31. J.S. Shang, P.G. Huang. Surface plasma actuators modeling for flow control. Progress in Aerospace Sciences. 2014;67:29–50. DOI: 10.1016/j.paerosci.2014.01.001
  32. 32. M. Abdollahzadeh, J.C. Pascoa, P.J. Oliveira. Implementation of the classical plasma-fluid model for simulation of dielectric barrier discharge (DBD) actuators in OpenFOAM. Computers & Fluids. 2016;108:77–99. DOI: 10.1016/j.compfluid.2016.01.012
  33. 33. S. L. Chernyshev , A. P. Kuryachii, D. A. Rusyanov, V. V. . Numerical modelling of plasma multi-actuator system. Blucher Mechanical Engineering Proceedings. 2014;1(1): 3026–3037. DOI: 10.5151/meceng-wccm2012–19129.
  34. 34. Likhanskii, M.N. Shneider, D.F. Opaits, R.B. Miles and S.O. Macheret. Numerical modelling of DBD plasma actuators and the induced air flow. In: 38th AIAA Plasmadynamics and Lasers Conference; 25 - 28 June 2007, Miami, FL. AIAA 2007–4533.
  35. 35. Hiroyuki Nishida, Taku Nonomura, and Takashi Abe. Characterization of Electrohydrodynamic Force on Dielectric-Barrier-Discharge Plasma Actuatorusing Fluid Simulation. International Journal of Mechanical, Aerospace, Industrial, Mechatronic and Manufacturing Engineering. 2012;6(11): 2324–2328. DOI:scholar.waset.org/1999.8/9312.
  36. 36. S. L. Chernyshev, A.P. Kuryachii, S.V. Manuilovich, D.A. Rusyanov, and V.V. Skvortsov. On the possibility of laminar flow control on a swept wing by means of plasma actuators. Progress in Flight Physics. 2015;7:169–190. DOI: 10.1051/eucass/201507169.
  37. 37. M. Abdollahzadeh, J. Páscoa, P. Oliveira. Numerical Modeling of Boundary Layer Control Using Dielectric Barrier Discharge. In: MEFTE 2012 - IV Fluid Mechanics, Thermodynamics and Energy National Conference; 28-29 May 2012; Lisbon, Portugal. 2012.
  38. 38. Y. Suzen, P. Huang, J.D. Jacob.Ashpis. Numerical simulations of plasma based flow control applications. In: 35th AIAA Fluid Dynamics Conference and Exhibit, Fluid Dynamics and Co-located Conferences; 6-9 June, 2005; Toronto, Canada. 2005. DOI: 10.2514/6.2005-4633
  39. 39. Y. Suzen, G. Huang, D. Ashpis. Numerical simulations of flow separation control in low pressure turbines using plasma actuators. In: 45th AIAA Aerospace Sciences Meeting and Exhibit; 8 - 11 January 2007; Reno, Nevada, AIAA-937-2007.
  40. 40. C. Enloe, T.E. McLaughlin, R.D. vanDyken, K. Kachner, E.J. Jumper, T. C Corke. Mechanisms and responses of a single dielectric barrier plasma actuator: plasma morphology. AIAA Journal . 2004; 42: 589-594.
  41. 41. D.M. Orlov. Modelling and simulation of single dielectric barrier discharge plasma actuators. [PhD Thesis]. 2006 D.M.Orlov.
  42. 42. B.E. Mertz, T.C. Corke, Time-dependent dielectric barrier discharge plasma actuator modeling. In: 47th AIAA Aerospace Sciences Meeting; Orlando, FL. AIAA Paper, vol.1083, 2009.
  43. 43. N. Benard, A. Debien, E. Moreau. Time-dependent volume force produced by a non-thermal plasma actuator from experimental velocity field. J. Phys. D: Appl. Phys; 46 (245201); 2013.
  44. 44. A. Kurz, S. Grundmann, C. Tropea, M. Forte, A. Seraudie, O. Vermeersch, D. Arnal, N. Goldin, R. King. Boundary Layer Transition Control using DBD Plasma Actuators. Journal Aerospace Lab. 2013; AL06-02.
  45. 45. A. Cristofolini, G. Neretti, C. Borghi. Experimental and numerical investigation on the surface charge distribution in a Dielectric Barrier Discharge fluid-dynamics plasma actuator. In: 43rd AIAA Plasmadynamics and Lasers Conference; 25-28 June, 2012; New Orleans, Louisian. 2012. p. AIAA-2012-3090.DOI: http://dx.doi.org/10.2514/6.2012-3090.
  46. 46. A. Criscione, B. Frohnapfel, C. Tropea. Preliminary Numerical Assessment of Turbulent Flow Control with Plasma Actuators. In: A. Dillmann, G. Heller, H. P. Kreplin, W. Nitsche, I. Peltzer. editors. New Results in Numerical and Experimental Fluid Mechanics VIII. Springer Berlin Heidelberg; 2013. p. 201-208. DOI: 10.1007/978-3-642-35680-3.
  47. 47. T. K. West IV. Numerical investigation of plasma actuator configurations for flow separation control at multiple angles of attack [ Master thesis]. 2012. Paper 5197.
  48. 48. G. Neretti, A. Cristofolini, and C.A. Borghi. Experimental investigation on a vectorized aerodynamic dielectric barrier discharge plasma actuator array. J. Appl. Physics. 2014; 115(163304). DOI: 10.1063/1.4873896.
  49. 49. N. Balcon, N. Benard, Y. Lagmich, J.P. Boeuf, G. Touchard, E. Moreau. Positive and negative sawtooth signals applied to a DBD plasma actuator influence on the electric wind. Journal of Electrostatics. 2009;67:140–145.
  50. 50. Benard N., E. Moreau. Role of the electric waveform supplying a dielectric barrier discharge plasma actuator. Applied Physics Letters. 2012; 100(9). ref. 193503.
  51. 51. M. Kotsonis and S. Ghaemi. Performance improvement of plasma actuators using asymmetric high voltage waveforms. J. Phys. D: Appl. Phys. 2012; 45 (045204). DOI: 10.1088/0022-3727/45/4/045204.
  52. 52. F.A. Dragonas, G. Neretti, P. Sanjeevikumar, and G. Grandi. High-Voltage High-Frequency Arbitrary Waveform Multilevel Generator for DBD Plasma Actuators. IEEE Transactions on Industry Applications. 2015; 51(4): 57-61. DOI:0.1109/TIA.2015.2409262.
  53. 53. D. Opaits, A. Likhanskii, G. Neretti, S. Zaidi, M. Shneider, R. Miles, and S. Macheret. Experimental Investigation on Dielectric Barrier Discharge Plasma Actuators Driven by Repetitive High-Voltage Nanosecond Pulses with DC or Low Frequency Sinusoidal bias. J. Appl. Physics. 2008; 104. DOI: 10.1063/1.2968251.
  54. 54. D. Opaits, G. Neretti, S. Zaidi, M. Shneider, R. Miles, A. Likhanskii, and S. Macheret. DBD Plasma Actuator Driven by a Combination of Low Frequency Bias Voltage and Nanosecond Pulses. In: 46th AIAA Aerospace Sciences Meeting and Exhibit; Reno, USA. AIAA-2008-1372; 2008.
  55. 55. R Dawson, J Little. Characterization of Nanosecond Pulse Driven Dielectric Barrier Discharge Plasma Actuators for Aerodynamic Flow Control. J. Appl. Physics. 2013; 113. DOI: 10.1063/1.4794507.
  56. 56. N. Benard, N. Zouzou, A. Claverie, J. Sotton, and E. Moreau. Optical visualization and electrical characterization of fast-rising pulsed dielectric barrier discharge for airflow control applications. J. Appl. Phys. 2012; 111 (033303). DOI: 10.1063/1.3682568.
  57. 57. R. Sosa, E. Arnaud, E. Memin, G. Artana. Study of the Flow Induced by a Sliding Discharge. IEEE Transactions on Dielectrics and Electrical Insulation. 2009; 16(2). P.305-311. DOI: 10.1109/TDEI.2009.4815157 .
  58. 58. E. Moreau, C. Louste, and G. Touchard,. Electric wind induced by sliding discharge in air at atmospheric pressure. J. Electrostat. 2008; 66(1-2), 107-114.
  59. 59. A. Santhanakrishnana and J. Jacob,. On Plasma Synthetic Jet Actuators. In: 44th AIAA Aerospace Science Meeting and Exhibit, Reno, NV, 2006. AIAA 2006-317. DOI: 10.2514/6.2006-317, (2006).
  60. 60. N. Bénard, J. Jolibois, E. Moreau, R. Sosa, G. Artana, and G. Touchard. Aerodynamic plasma actuators: A directional micro-jet device. J Thin Solid Films. 2008; 516; 6660-6667. DOI: 10.1016/j.tsf.2007.11.039.
  61. 61. S.H. Kim,W. Hong, and C. Kim. Separation Control Mechanism of Airfoil Using Synthetic Jet. J. Mechanical Science and Technology. 2007; 21, 1367-1375.
  62. 62. N. Bénard, J. Jolibois, G. Touchard, and E. Moreau. A Directional Plasma-Jet Device Generated by Double DBD Actuators - An Active Vortex Generator for Aerodynamic Flow Control. In: 4th Flow Control Conference, Seattle, Washington. 2008. AIAA 2008-3763.
  63. 63. R Joussot, A Leroy, R Weber, H Rabat, S Loyer and D Hong. Plasma morphology and induced airflow characterization of a DBD actuator with serrated electrode. J. Phys. D: Appl. Phys. 2013; 46(125204); 12pp. DOI: 10.1088/0022-3727/46/12/125204.
  64. 64. R J Durscher and S Roy. Three-dimensional flow measurements induced from serpentine plasma actuators in quiescent air. J. Phys. D: Appl. Phys. 2012; 45 (035202);9pp. DOI:10.1088/0022-3727/45/3/035202.
  65. 65. Y.E. Akansu, F. Karakaya, and A. Şanlısoy. Active Control of Flow around NACA 0015 Airfoil by Using DBD Plasma Actuator. EPJ Web of Conferences. 2013; 45(01008): 7pp. DOI: 10.1051/epjconf/20134501008.
  66. 66. C.A. Borghi, A. Cristofolini, G, Neretti, P. Seri, A. Talamelli, and A. Rossetti. Wind Tunnel Experiments on a NACA0015 Airfoil Equipped with Vectorizable
  67. 67. M.L. Post and T.C. Corke. Separation Control on High Angle of Attack Airfoil Using Plasma Actuators. AIAA Journal. 2004; 42: pp. 2177-2184.
  68. 68. C. He, T.C. Corke, and M.P. Patel. Plasma Flaps and Slats: An Application of Weakly Ionized Plasma Actuators. J. Aircraft. 2009; 46; pp. 864-873. DOI: 10.2514/1.38232.
  69. 69. D.P. Rizzetta and M.R. Visbal. Numerical Investigation of Plasma-Based Control for Low-Reynolds-Number Airfoil Flows. AIAA Journal. 2011; 49: pp. 411-425. DOI: 10.2514/1.J050755.
  70. 70. Mark Riherd and Subrata Roy. Serpentine geometry plasma actuators for flow control. J. of Appl. Physics D. 2013; 114(083303). DOI: 10.1063/1.4818622
  71. 71. K. Inaoka, T. Mori, M. Yamaguchi, and M. Senda. Feedback flow control of a low-Re airfoil by flap actuators. Journal of Fluids and Structures. 2015; 58: 319-330. DOI: 10.1016/j.jfluidstructs.2015.08.011.
  72. 72. T. C. Corke, P.O. Bowles, C. He, and E. Matlis. Sensing and control of flow separation using plasma actuators. Phil. Trans. R. Soc. A. 2011; 369: 1459-1475. DOI: 10.1098/rsta.2010.0356
  73. 73. Maden, R. Maduta, S. Jakirlic, S. Grundmann, C. Tropea, and J.K. Eaton. Plasma-Based
  74. 74. S. Grundmann, E. L. Sayles, and J. K. Eaton. Sensitivity of an asymmetric 3D diffuser to plasma-actuator induced inlet condition perturbations. Experiments in Fluids. 2010; 50(1):217–23. DOI: 10.1007/s00348-010-0922-0.
  75. 75. Potočar, E., Širok, B., Hočevar, M., and Eberlinc, M. Control of Separation Flow over a Wind Turbine Blade with Plasma Actuators. Journal of Mechanical Engineering. 2012; 58(1): 37-45. DOI: 10.5545/sv-jme.2011.016.
  76. 76. Yong Li, Xin Zhang, and Xun Huang. The use of plasma actuators for bluff body broadband noise control. Experiments in Fluids. 2010; 49(2); 367-377. DOI: 10.1007/s00348-009-0806-3.
  77. 77. F. O. Thomas, A. Kozlov, and T. C. Corke. Plasma Actuators for Cylinder Flow Control and Noise Reduction. AIAA Journal. 2008; 46(8): pp. 1921-1931.
  78. 78. Xun Huang, and Xin Zhang. Plasma actuators for noise control. International Journal of Aeroacoustics. 2010; 9(4 & 5): pages 679-704.
  79. 79. C. L. Kelley, P. O. Bowles, J. Cooney, C. He, T.C. Corke, B. A. Osborne, J.S. Silkey, and J. Zehnle. Leading-Edge Separation Control Using Alternating-Current and Nanosecond-Pulse Plasma Actuators. AIAA Journal. 2014; 52(9): 1871-1884.doi: 10.2514/1.J052708
  80. 80. Y. Bouremela, J.M. Lia, Z. Zhaob, and M. Debiasia. Effects of AC Dielectric Barrier Discharge Plasma Actuator Location on Flow Separation and Airfoil Performance. Procedia Engineering. 2013; 67: 270-278. DOI: 10.1016/j.proeng.2013.12.026.
  81. 81. Hikaru Aono, Taku Nonomura, Aiko Yakeno, Kozo Fujii and Koichi Okada. Plasma Flow Control Simulation of an Airfoil of Wind Turbine at an Intermediate Reynolds Number. In: 8th Symposium on Flow Manipulation and Active Control: Theory, Experiments and Implementation; July 7-11, 2013; Nevada, USA, . 2013. DOI: 978-0-7918-5555-3.
  82. 82. Miguel Visbal, Datta Gaitonde, and Subrata Roy. Control of Transitional and Turbulent Flows Using Plasma-Based Actuators. In: 36th AIAA Fluid Dynamics Conference and Exhibit, Fluid Dynamics and Co-located Conferences; 5-8 June, 2006; San Francisco, California. 2006. AIAA-2006-3230. DOI: 10.2514/6.2006-3230 .
  83. 83. A. N. M. M. I. Mukut, H. Mizunuma, O. Hiromichi, T. Segawa. Winglet type dielectric barrier discharge plasma actuators: performance characterization and numerical comparison. Procedia Engineering. 2015; 105: 250-257. DOI: http://dx.doi.org/10.1016/j.proeng.2015.05.104.
  84. 84. Xiaohu Zhao, Yinghong LI, Yun Wu, Tao Zhu, Yiwen Li. Numerical Investigation of Flow Separation Control on a Highly Loaded Compressor Cascade by Plasma Aerodynamic Actuation . 2012;25(3):349–360. DOI: 10.1016/S1000-9361(11)60396-8
  85. 85. Philipp C. Dörr, Markus, J. Kloker. Transition Control in a Three-dimensional Boundary-layer Flow Using Plasma Actuators. Procedia IUTAM. 2015;14:469–478. DOI: 10.1016/j.piutam.2015.03.075.
  86. 86. I. Maden, R. Maduta, J. Kriegseis, S. Jakirlič, C. Schwarz, S. Grundmann, C. Tropea. Experimental and computational study of the flow induced by a plasma actuator. International Journal of Heat and Fluid Flow. 2013;41:80–89. DOI: 10.1016/j.ijheatfluidflow.2013.02.013
  87. 87. E. Peersa, Z. Ma, X. Huang . A numerical model of plasma effects in flow control. Physics Letters A. 2010;374(13-14):1501–1504. DOI: 10.1016/j.physleta.2009.08.046
  88. 88. A. Kurz, S. Grundmann, C. Tropea, M. Forte, A. Seraudie, O. Vermeersch, D. Arnal, N. Goldin, R. King. Boundary Layer Transition Control using DBD Plasma Actuators. Journal Aerospace Lab. 2013; 6: AL-06-02. ISSN: 2107-6596.

Written By

Gabriele Neretti

Submitted: 30 September 2015 Reviewed: 25 February 2016 Published: 08 September 2016