Detailed simulation of turbulent flow within a suction and oscillatory blowing fluidic actuator

Similar documents
Development of Simplified Boundary Condition of SaOB Actuator Based on High-Fidelity CFD Simulations

DNS, LES, and wall-modeled LES of separating flow over periodic hills

Suction and Oscillatory Blowing Actuator

STUDY OF THREE-DIMENSIONAL SYNTHETIC JET FLOWFIELDS USING DIRECT NUMERICAL SIMULATION.

Large eddy simulation of turbulent flow over a backward-facing step: effect of inflow conditions

Available online at ScienceDirect. Procedia Engineering 79 (2014 ) 49 54

Boundary-Layer Theory

Turbulent drag reduction by streamwise traveling waves

Computation of turbulent Prandtl number for mixed convection around a heated cylinder

An evaluation of a conservative fourth order DNS code in turbulent channel flow

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration

Large-eddy simulations for wind turbine blade: rotational augmentation and dynamic stall

Direct Numerical Simulation of Jet Actuators for Boundary Layer Control

SELF-SUSTAINED OSCILLATIONS AND BIFURCATIONS OF MIXED CONVECTION IN A MULTIPLE VENTILATED ENCLOSURE

Prediction of unsteady heat transfer from a cylinder in crossflow

Numerical simulation of lift enhancement on a NACA 0015 airfoil using ZNMF jets

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

Predictions of Aero-Optical Distortions Using LES with Wall Modeling

Numerical investigation of swirl flow inside a supersonic nozzle

Active Control of Separated Cascade Flow

Table of Contents. Foreword... xiii. Preface... xv

Toward Simple Boundary Condition Representations of Zero-Net Mass-Flux Actuators in Grazing Flow

Numerical simulations of the edge tone

A combined application of the integral wall model and the rough wall rescaling-recycling method

Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows

Numerical Simulation of Flow Separation Control using Multiple DBD Plasma Actuators

Principles of Convection

Simulating Drag Crisis for a Sphere Using Skin Friction Boundary Conditions


CONTROL OF FLOW SEPARATION BY TIME PERIODIC ELECTROMAGNETIC FORCES

Application of the immersed boundary method to simulate flows inside and outside the nozzles

FEDSM COMPUTATIONAL AEROACOUSTIC ANALYSIS OF OVEREXPANDED SUPERSONIC JET IMPINGEMENT ON A FLAT PLATE WITH/WITHOUT HOLE

Periodic planes v i+1 Top wall u i. Inlet. U m y. Jet hole. Figure 2. Schematic of computational domain.

Force analysis of underwater object with supercavitation evolution

Numerical Study of Jet Plume Instability from an Overexpanded Nozzle

Validation 3. Laminar Flow Around a Circular Cylinder

arxiv: v1 [physics.flu-dyn] 11 Oct 2012

A dynamic global-coefficient subgrid-scale eddy-viscosity model for large-eddy simulation in complex geometries

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION

GENERALISATION OF THE TWO-SCALE MOMENTUM THEORY FOR COUPLED WIND TURBINE/FARM OPTIMISATION

ACTIVE SEPARATION CONTROL PROCESS OVER A SHARP EDGE RAMP

Basic Features of the Fluid Dynamics Simulation Software FrontFlow/Blue

DIRECT NUMERICAL SIMULATION OF SPATIALLY DEVELOPING TURBULENT BOUNDARY LAYER FOR SKIN FRICTION DRAG REDUCTION BY WALL SURFACE-HEATING OR COOLING

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

There are no simple turbulent flows

Wall turbulence with arbitrary mean velocity profiles

Canada. CFD-Experiments Integration in the Evaluation of Six Turbulence Models for Supersonic Ejectors Modeling

SEPARATION CONTROL BY SYNTHETIC JET ACTUATOR IN A STRAIGHT BLADE CASCADE

Basic Fluid Mechanics

Shock/boundary layer interactions

Turbulent Boundary Layers & Turbulence Models. Lecture 09

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

Fluid Dynamics Exercises and questions for the course

2 Navier-Stokes Equations

Predicting natural transition using large eddy simulation

LES of synthetic jets in boundary layer with laminar separation caused by adverse pressure gradient

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

Active Control of Turbulence and Fluid- Structure Interactions

Numerical simulation of wave breaking in turbulent two-phase Couette flow

WALL RESOLUTION STUDY FOR DIRECT NUMERICAL SIMULATION OF TURBULENT CHANNEL FLOW USING A MULTIDOMAIN CHEBYSHEV GRID

Large Eddy Simulation of Three-Stream Jets

DYNAMIC SEPARATION CONTROL IN A LOW-SPEED ASYMMETRIC DIFFUSER WITH VARYING DOWNSTREAM BOUNDARY CONDITION

SIMULATION OF THREE-DIMENSIONAL INCOMPRESSIBLE CAVITY FLOWS

Local correlations for flap gap oscillatory blowing active flow control technology

Numerical Simulation of Rocket Engine Internal Flows

Experimental and Numerical Investigation of Flow over a Cylinder at Reynolds Number 10 5

Develpment of NSCBC for compressible Navier-Stokes equations in OpenFOAM : Subsonic Non-Reflecting Outflow

Elliptic Trailing Edge for a High Subsonic Turbine Cascade

EXPERIMENTS OF CLOSED-LOOP FLOW CONTROL FOR LAMINAR BOUNDARY LAYERS

Implicit Large Eddy Simulation of Transitional Flow over a SD7003 Wing Using High-order Spectral Difference Method

Natural convection adjacent to a sidewall with three fins in a differentially heated cavity

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM

Computational Investigations of High-Speed Dual-Stream Jets

Computational Fluid Dynamics Analysis of Jets with Internal Forced Mixers

SIMULATION OF PRECESSION IN AXISYMMETRIC SUDDEN EXPANSION FLOWS

Experimental Study on Flow Control Characteristics of Synthetic Jets over a Blended Wing Body Configuration

The effect of geometric parameters on the head loss factor in headers

Defense Technical Information Center Compilation Part Notice

Study of Forced and Free convection in Lid driven cavity problem

Evolution of the pdf of a high Schmidt number passive scalar in a plane wake

Compressible turbulent channel flow with impedance boundary conditions

Flow analysis in centrifugal compressor vaneless diffusers

Self-Excited Vibration in Hydraulic Ball Check Valve

Parash Moni Thakur. Gopal Ch. Hazarika

A Comparative Analysis of Turbulent Pipe Flow Using k And k Models

Effects of Free-Stream Vorticity on the Blasius Boundary Layer

On the feasibility of merging LES with RANS for the near-wall region of attached turbulent flows

COMPUTATIONAL SIMULATION OF THE FLOW PAST AN AIRFOIL FOR AN UNMANNED AERIAL VEHICLE

Turbulent eddies in the RANS/LES transition region

Turbulent boundary layer

Design and Aerodynamic Characterization of a Synthetic Jet for Boundary Layer Control

Given the water behaves as shown above, which direction will the cylinder rotate?

Numerical simulations of oblique shock/boundary-layer interaction at a high Reynolds number

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

The JHU Turbulence Databases (JHTDB)

Heat Transfer from An Impingement Jet onto A Heated Half-Prolate Spheroid Attached to A Heated Flat Plate

Pulsation Amplitude Influence on Free Shear Layer of a Vertical Pulsating Jet in Laminar Regime

AA210A Fundamentals of Compressible Flow. Chapter 5 -The conservation equations

Application of Compact Schemes to Large Eddy Simulation of Turbulent Jets

Modeling of Wall Heat Transfer and Flame/Wall Interaction A Flamelet Model with Heat-Loss Effects

Transcription:

Center for Turbulence Research Annual Research Briefs 14 9 Detailed simulation of turbulent flow within a suction and oscillatory blowing fluidic actuator By J. Kim AND P. Moin 1. Motivation and objectives Fluidic actuators have been widely used for several decades to control turbulent flows. Although their application and objectives vary depending on the particular configurations of turbulent flows and actuator types, the basic mechanism of fluidic actuators is the same: adding or removing fluid momentum to perturb, at least, the local state of the flow so that an integral quantity of interest (for example, skin friction, drag coefficients, overall sound pressure level) is improved. Several types of fluidic actuators can be found with their mechanisms based upon, for example, steady suction, blowing and suction with zero-net mass flux, resonance, or synthetic jet (Gad-el Hak ; Greenblatt & Wygnanski ; Collis et al. 4; Cattafesta III & Sheplak 11; Mohseni & Mittal 14). Recently, a novel fluidic actuator using steady suction and oscillatory blowing was developed for the active control of high-speed turbulent flows (Arwatz et al. 8). The concepts of suction and oscillatory blowing are integrated into a single active control device for higher control authority. The actuation is based upon a self-sustained mechanism of unsteady flows and does not require any moving parts. The output is controlled primarily by its pressure source and the feedback tube (discussed in Section 2), and no additional input is needed. This new fluidic actuator demonstrates a robust and effective performance for delaying boundary-layer separation and bluff-body drag reduction. The internal flow in this actuator has been subject to detailed examination in an effort to quantify the actuator s behavior as well as to develop a predictive model for the actuation (Arwatz et al. 8; Wassermann et al. 13). Arad (14) conducted both unsteady Reynolds Averaged Navier Stokes (URANS) and incompressible large eddy simulation (LES) using the same grid. It was argued that URANS could predict the oscillatory jet based upon the observation that a key contributing factor to jet oscillation is the Coandă effect, which is not strongly turbulent motion. Although LES was able to reproduce some of the important features of the oscillatory jet within the actuator, the oscillatory behavior was not quite clear for the URANS results. Both calculations demonstrated some discrepancy with the experimental measurements. Several inconsistencies were pointed out for LES, such as using an incompressible flow solver, neglecting the suction box, and prescribing steady suction as a boundary condition. In this study, a high-fidelity simulation of the internal jet flow within the fluidic actuator is performed. A fully compressible LES solver is used and a complete geometry of the actuator is simulated. In Section 2, the Suction and Oscillatory Blowing actuator is described. Computational conditions are discussed in Section 3, followed by simulation results in Section 4. The report concludes with a summary and outlook for future work in Section 5.

2 Kim & Moin x Plenum y z Switching valve Actuator outlets Ejector Splitter Nozzle (actuator inlet) Feedback tube Suction box Figure 1. The SaOB fluidic actuator. The block arrow indicates the main flow direction. 2. Suction and Oscillatory Blowing (SaOB) fluidic actuator In this section, the SaOB fluidic actuator and its operating mechanism are briefly explained. A more detailed description can be found in Arwatz et al. (8). The schematic diagram for the full-geometry SaOB fluidic actuator is shown in Figure 1. Throughout this report, x is the streamwise direction, y is the transverse direction, and z is the spanwise direction. The center-plane corresponds to y = around which the actuator geometry is symmetric, except for the feedback tube. A pressure source such as a stagnation chamber is connected to the converging nozzle, and a round jet is injected into the ejector with a square cross section. The flow direction is given by the block arrow in Figure 1. The suction box connected to the atmospheric or lower-pressure air supplies an appropriate amount of entrainment through the holes on the suction-box surface. The suction box is connected to the ejector via two suction ports. The amount of suction is dynamically determined by the pressure condition within the ejector, which determines the total flow rate of the actuator. The round jet develops within the ejector and enters the switching valve after a mildly converging section. The switching valve has two actuator outlets (top and bottom for y > and y <, respectively) downstream of the splitter. The S-shaped feedback tube connects the upper and lower pressure ports and is essential for the self-sustained operation of the actuator. Under certain operating conditions, the plane jet within the switching valve oscillates in time. First, the jet loses symmetry with respect to y =, presumably due to the amplified unstable modes of the plane jet. Then, it deflects to one side and eventually attaches to the corresponding wall in y direction. A transverse pressure gradient caused by the Coandă effect (Tritton 1986) pushes the jet close to the wall and the jet eventually attaches to the wall. Without the feedback tube, the wall-attached jet is stable in time and does not oscillate. However, the presence of the feedback tube enables momentum transfer from one side of the pressure ports to the other. The momentum is driven by the pressure difference between the two pressure ports. The pressure port on the side where the jet is deflected has lower pressure due to the local acceleration caused by the Coandă effect. Subsequently, pressure on the other side of the pressure port becomes relatively higher and momentum is transferred to the lower-pressure port. The resulting transverse pressure gradient competes with the Coandă force stabilzing the jet near the wall and eventually detaches the jet. Once detached from the wall and crossingy =, the jet moves quickly to the other side due to the cooperative effect between the transverse pressure gradient and the Coandă force. The wall-attached jet detaches in the same way

Suction and Oscillatory Blowing fluidic actuator 211 and re-attaches to the original side of the wall, thus completing a single oscillation cycle. The oscillation is repeated in a self-sustained manner at a certain frequency determined primarily by the length and cross-sectional area of the feedback tube, temperature within the tube, and the nozzle inlet pressure (Viets 1975; Raman et al. 1994). The SaOB fluidic actuator is robust and effective in providing high-speed, bi-stable oscillatory blowing at a prescribed frequency by the self-sustained mechanism. An important characteristic of the actuator is that none of its components move to produce the oscillatory blowing and no additional input is needed to sustain the oscillation; the main control input is the nozzle inlet pressure, which determines the overall flow rate of the actuator. Another advantage is the amplification of the flow rate caused by the low-pressure cavity formed within the switching valve when the jet is attached to one side of the walls. The low-pressure zone entrains fluid outside of the actuator outlet and increases the amount of blowing. Its high control authority is promising in actively controlling turbulent flows. Wilson et al. (13) applied an array of 28 SaOB actuators synchronized in phase to delay boundary-layer separation and reduce the drag of an axisymmetric body. Its effectiveness as an active flow control device is demonstrated experimentally for canonical flow configurations (Wilson et al. 13; Shtendel & Seifert 14). 3. Computational conditions The high-fidelity, computational fluid dynamics code, CharLES, is used to solve the fully compressible Navier Stokes equations. The governing equations are non-dimensionalized by the nozzle radius r j = D j /2, the ambient speed of sound, density ρ, temperature T, and molecular viscosity µ, where the subscript denotes the ambient state. The ideal gas is assumed with γ = c p /c v = 1.4, where c p and c v are the specific heats at a constant pressure and volume, respectively. The viscosity is assumed to be a function of temperature only and to follow the power-law relation, µ/µ = (T/T ) n where n =.76. The Reynolds number based on the nozzle-exit condition (denoted by the subscript j) is Re j = ρ j u j D j /µ j = 64 and the Prandtl number is assumed constant as Pr =.7. The subgrid-scale model of Vreman (4) is used with the model constant c =.7 and the constant turbulent Prandtl number Pr t =.9. The governing equations are discretized in space using the cell-based finite volume formulation. The solutions are time advanced using the standard third-order Runge Kutta method at a constant (non-dimensional) time-step size of t =.75, which results in the Courant Friedrichs Lewy number of approximately 1.. The spatial discretization is non-dissipative and formally second-order accurate on arbitrary unstructured grids. In addition, the convective fluxes are combined, depending on local grid quality (for example, element skewness), with fluxes computed by an HLLC-upwind discretization. Khalighi et al. (11) provide more detailed discussions on the spatial discretization. A base grid of 19 million unstructured control volumes is generated and subsequently refined using Adapt, the grid-adapting tool in the CharLES suites. For computational accuracy and efficiency, only hexahedral elements are used. The total number of control volumes for the adapted grid is 37 million. A cross-section of the computational grid is shown in Figure 2. The nozzle is supplied with air at 24 psi, and its total temperature is the same as the ambient temperature. The Mach number at the nozzle exit is M j = u j /c j =.87. The http://www.cascadetechnologies.com/pdf/charles.pdf

212 Kim & Moin Figure 2. The computational grid for the ejector and switching valve. suction holes and the plenum outlet are at the ambient condition. The no-slip solid wall boundaries are modeled isothermally at the ambient temperature. The simulation is performed in the ALCF Mira cluster using 8,192 cores. There are approximately 4,5 control volumes per core, and the code scales very well up to this number of cores. 4. Results 4.1. Self-sustained oscillation of the ejector jet The flow within the SaOB actuator is time advanced until initial transients are swept away and the jet in the switching valve establishes a bi-stable oscillation. To initiate an oscillatory state, steady suction is applied on the upper wall of the switching valve (y > ) until the jet attaches to it, after which the wall is switched back to a noslip and no-penetration solid wall. The oscillation frequency f = 1/T is estimated based upon the time history of the volume flow rate on the actuator outlets shown in Figure 3. The current simulation predicts the switching frequency to be 243 Hz, while the experiment measures 211 Hz under the corresponding inlet pressure. The oscillation frequency can also be estimated by calculating the volume flow rates on the pressure ports, as shown in Figure 4(a). The volume flow rates can be accurately modeled by a sinusoidal function with a constant period superimposed by high-frequency fluctuations. The lower and upper pressure ports are synchronous and in-phase in time, consistent with the measurements (Arwatz et al. 8). Note that the measured frequency in the experiments is not constant but fluctuates over the oscillation periods by about Hz, presumably due to strong turbulent fluctuations within the switching valve. The first column of Figure 5 shows the instantaneous velocity contours on z = at different phases of the oscillation. The jet demonstrates an apparent switching between the upper and lower sections of the switching valve. Large-scale instability wave-like structures are clearly seen. Note that t in Figure 5 is same as t in Figure 3. In the second column, instantaneous velocity magnitudes on the actuator outlets are shown at each phase of the oscillation. The outlet to which the jet is deflected has the maximum velocity of.28. On the other side, the instantaneous velocity is approximately uniform in the transverse direction with a magnitude of % to 4% of the maximum velocity.

Suction and Oscillatory Blowing fluidic actuator 213 6 5 Top Bottom Q/(c r 2 j) 4 3 2 1..2.4.6.8 1. 1.2 t /T Figure 3. Volume flow rates on the top and bottom sides of the actuator outlets. The horizontal axis is the relative time with its reference defined in Figure 5(a). The oscillation period is denoted by T..4.2 Top Bottom.2sin(2πt /T +7π/8) 2.4 2.2 Top Bottom Q/(c r 2 j). Q/(c r 2 j) 2. -.2 1.8 -.4 1.6..2.4.6.8 1. 1.2 t /T (a)..2.4.6.8 1. 1.2 t /T (b) Figure 4. Volume flow rates on the top and bottom sides of (a) the pressure ports and (b) the suction ports. The horizontal axis is the relative time with its reference defined in Figure 5(a). The oscillation period is denoted by T. 4.2. Statistics The solutions are time averaged and statistics are compared with the corresponding experiment. Due to the substantially low non-dimensional frequency (f r j / = 9.8 4 ), statistical samples are collected only for a single oscillation period; additional simulations will be performed to obtain converged statistics for several oscillation periods. The quantity of particular interest for this actuator is the velocity at the actuator outlet. In Figure 6(a), contours of time-averaged streamwise velocity are shown. The locations of maximum velocity lie close to the splitter wall. Figure 6(b) shows time-averaged velocities at the spanwise centerline on the actuator outlet. For plotting purposes, the sign of the y-velocity profile in the bottom side of the outlet is switched for comparison with the profile in the upper side. The profiles are fairly symmetric and the maximum streamwise velocity is.16, corresponding to 56 m/s. The experiment reports a maximum velocity of 44 m/s, a 27% error compared to the current prediction.

214 Kim & Moin u / - - (a) t /T = (b) t /T = u / - - (c) t /T = 1/4 (d) t /T = 1/4 u / - - (e) t /T = 1/2 (f) t /T = 1/2 Figure 5. For the caption, see the next page.

Suction and Oscillatory Blowing fluidic actuator 215 u / - - (g) t /T = 3/4 (h) t /T = 3/4 u / - - (i) t /T = 1 (j) t /T = 1 Figure 5. Instantaneous contours of streamwise velocity (left column) and transverse profiles of velocity magnitude (right column). The suction box shown in Figure 1 entrains the ambient air into the ejector. The round jet within the ejector has static pressure lower than the ambient pressure. This causes a pressure difference between the ejector and the suction box connected to the ambient air. Thus, air is drawn into the suction box and flows into the ejector. In the previous computational study(arad 14), this entrainment was modeled as a boundary condition of a constant volume flow rate, which matches the corresponding experimental data in a time-averaged sense. The volume flow rates through the suction ports are shown in Figure 4(b) and the mean volume flow rate is 2.6 3 m 3 /s. This is an approximately 25% underprediction of the experimental measurement. The reduced entrainment is a consequence of a lower pressure difference between the ambient air and the ejector jet. The static pressure within the suction box is nearly uniform (within.2%) as the ambient pressure. This suggests that the pressure of the ejector jet is overpredicted. Another possibility is that turbulent mixing of the ejector jet is not as strong as in the experiment, resulting in less ambient air entrainment. For the present simulation, the spatial resolution within the ejector may have a strong

216 Kim & Moin ū..15 Top Bottom ū. ū/ v/.5 y/r j z/r j. 5 15 y/r j (a) (b) Figure 6. Time-averaged velocity on the actuator outlet: (a) streamwise velocity contour and (b) velocity profiles at z =.2. impact on the development of the turbulent jet. Currently, the ejector grid is further refined and additional simulation is being performed to assess this hypothesis. In any case, the reduced entrainment is associated with a reduced growth rate of the jet in the streamwise direction. Consequently, the jet centerline velocity decays more slowly, and thus the velocity at the ejector exit becomes higher. The increased velocity at the ejector exit appears to result in higher frequency jet oscillations in the switching valve, which is consistent with the current observation and the experiment (Arwatz et al. 8). 5. Summary and future work Turbulent flow within the Suction and Oscillatory Blowing (SaOB) actuator is simulated. The SaOB fluidic actuator is developed to provide high-momentum oscillatory blowing for active flow control, such as separation delay of the high-speed boundary layer over airfoils. Its operating mechanism is based upon a balance between the transverse pressure gradient generated by the oscillating jet and the Coandă force, and it does not involve any moving parts. Its control authority has been demonstrated for several canonical flow configurations. Providing detailed and reliable description of the unsteady internal flow is important for a better understanding of the physical mechanisms and eventually improving the actuator design. Large-eddy simulation technique is applied to compute the compressible turbulent flow within the actuator. The actuator geometry is represented using an adapted unstructured grid with 37 million control volumes. The simulation predicts the time-dependent switching behavior of the turbulent jet. The oscillatory motion repeats in a self-sustained manner at a frequency close to the experimental measurement. Despite some quantitative discrepancies with the experiment, the preliminary results are encouraging. Simulations with more increased grid resolution using 28 million control volumes are underway. Additional simulations will be performed using inlet pressuresof5,,and psi to parametricallyassessthe fidelity ofthe current prediction.

Acknowledgments Suction and Oscillatory Blowing fluidic actuator 217 The primary computing resources were provided by the Argonne National Laboratory through the ASCR Leadership Computing Challenge. The authors also acknowledge National Science Foundation for supporting the Certainty cluster at the Center for Turbulence Research. Financial support from NASA and the Boeing Company is gratefully acknowledged. The authors would like to thank Prof. Avraham Seifert for providing the actuator geometry and engaging in valuable discussions. The first author is grateful to Dr. George Ilhwan Park, Dr. Yee Chee See, Ik Jang, and Brian Pierce for help in using the CharLES code and for useful discussions. REFERENCES Arad, E. 14 The appropriate level of simulation for a fluidic oscillator AFC device (private communication) pp. 1 12. Arwatz, G., Fono, I. & Seifert, A. 8 Suction and oscillatory blowing actuator modeling and validation. AIAA J. 46, 17 1117. Cattafesta III, L. N. & Sheplak, M. 11 Actuators for active flow control. Annu. Rev. Fluid Mech. 43, 247 272. Collis, S. S., Joslin, R. D., Seifert, A. & Theofilis, V. 4 Issues in active flow control: theory, control, simulation, and experiment. Prog. Aerosp. Sci. 4, 237 289. Greenblatt, D. & Wygnanski, I. J. The control of flow separation by periodic excitation. Prog. Aerosp. Sci. 36, 487 545. Gad-el Hak, M. Flow Control: Passive, Active, and Reactive Flow Management. Cambridge University Press. Khalighi, Y., Nichols, J. W., Lele, S. K., Ham, F. & Moin, P. 11Unstructured large eddy simulation for prediction of noise issued from turbulent jets in various configurations. AIAA Paper 11 2886. Mohseni, K. & Mittal, R. 14 Synthetic Jets: Fundamentals and Applications. CRC Press. Raman, G., Rice, E. J. & Cornelius, D. M. 1994 Evaluation of flip-flop jet nozzles for use as practical excitation devices. J. Fluids Eng. 116, 58 515. Shtendel, T. & Seifert, A. 14 Three-dimensional aspects of cylinder drag reduction by suction and oscillatory blowing. Int. J. Heat Fluid Fl. 45, 9 127. Tritton, D. J. 1986 Physical Fluid Dynamics. Oxford Science Publications. Viets, H. 1975 Flip-flop jet nozzle. AIAA J. 13, 1375 1379. Vreman, A. W. 4 An eddy-viscosity subgrid-scale model for turbulent shear flow: Algebraic theory and applications. Phys. Fluids 16, 367 3681. Wassermann, F., Hecker, D., Jung, B., Markl, M., Seifert, A. & Grundmann, S. 13 Phase-locked 3D3C-MRV measurements in a bi-stable fluidic oscillator. Exp. Fluids 54, 1 15. Wilson, J., Schatzman, D., Arad, E., Seifert, A. & Shtendel, T. 13 Suction and pulsed-blowing flow control applied to an axisymmetric body. AIAA J. 51, 2432 2446.