You are on page 1of 534

Fundamentals and Applications

of Modern Flow Control

Edited by
Ronald D. Joslin
Office of Naval Research
Arlington, Virginia

Daniel N. Miller
Lockheed Martin Aeronautics Company
Fort Worth, Texas

Volume 231
PROGRESS IN
ASTRONAUTICS AND AERONAUTICS

Frank K. Lu, Editor-in-Chief


University of Texas at Arlington
Arlington, Texas

Published by the
American Institute of Aeronautics and Astronautics, Inc.
1801 Alexander Bell Drive, Reston, Virginia 20191-4344
The cover images are declared a work of the U.S. Government. The shadowgraph image showing jet
vectoring is courtesy of Jeffrey D. Flamm, NASA Langley Research Center, and presented in AIAA
Paper 2005-3503. The XV-15 tilt rotor image that was used in a modern flow control flight experiment
is courtesy of NASA Dryden Flight Research Center (EC80-13848).

American Institute of Aeronautics and Astronautics, Inc., Reston, Virginia


1 2 3 4 5
Copyright © 2009 by the American Institute of Aeronautics and Astronautics, Inc. Printed in the United
States of America. All rights reserved. Reproduction or translation of any part of this work beyond that
permitted by Sections 107 and 108 of the U.S. Copyright Law without the permission of the copyright
owner is unlawful. The code following this statement indicates the copyright owner’s consent that
copies of articles in this volume may be made for personal or internal use on condition that the copier
pay the copy fee ($2.50) plus the per-page fee ($0.50) through the Copyright Clearance Center Inc., 222
Rosewood Drive, Danvers, Massachasetts 01923. This consent does not extend to other kinds of copy-
ing, for which permission requests should be addressed to the publisher. Users should employ the
following code when reporting copying from the volume to the Copyright Clearence Center

978-1-56347-983-0/09-$2.50+0.50

Data and information appearing in this book are for informational purposes only. AIAA is not responsi-
ble for any injury or damage resulting from use or reliance, nor does AIAA warrant that use or reliance
will be free from privately owned rights.
Print ISBN 978-1-56347-983-0
e-book ISBN 978-1-56347-988-5
Acronyms and Abbreviations

2-D = two-dimensional
3-D = three-dimensional
AC = alternating current
ACS = active control system
AVF = acoustic-vortex-flame
ADVINT = adaptive flow control vehicle integrated technologies
AFC = active flow control
AIP = aerodynamic interface plane
AM = amplitude modulation
AoA = angle of attack
ASC = active separation control
BLC = boundary-layer control
BM = burst modulation
BTW = Boeing tilt-wing
BVI = blade–vortex interaction
CCD = charge-coupled device
CF = crossflow
CFD = computational fluid dynamics
CI = combustion instability
CIACS = combustion instability active control system
CLAFC = closed-loop active flow control
CO = carbon monoxide
DARPA = Defense Advanced Research Projects Agency
DBD = dielectric barrier discharge
DC = direct current
DES = detached eddy simulation
DNS = direct numerical simulation
DOE = design of experiments
DRE = distributed roughness element
DSV = dynamic stall vortex
DTN = dual throat nozzle
EET = energy-efficient transport
EHD = electrohydrodynamic
EPNdB = effective perceived noise level, dB
FSM = flow-simulation methodology
FST = freestream turbulence
FSTI = freestream turbulence intensity
FW-H = Ffowcs–Williams and Hawkings
HHI = higher harmonic control
HIFEX = high-frequency excitation (actuators)
HPT = high-pressure turbine

xix
xx

HSI = high-speed impulsive


IB = immersed boundary
ILES = implicit large-eddy simulation
K-H = Kelvin–Helmholtz instability
LBO = lean blowout
LBOACS = lean blowout active control system
LDI = lean direct injection
LE = leading edge
LEM = lumped element model
LES = large-eddy simulation
LFC = laminar flow control
LNSE = linear Navier–Stokes equations
LPP = lean prevaporized premixed
LPT = low-pressure turbine
LQG = linear–quadratic–Gaussian
LQR = linear–quadratic regulator
LTI = linear time-invariant
MATV = multi-axis thrust vectoring
MEMS = micro-electromechanical systems
MIMO = multi-input, multi-output
MIT = Massachusetts Institute of Technology
NACA = National Advisory Committee for Aeronautics
N-S = Navier–Stokes (equations)
NO = nitric oxide
NO2 = nitrogen dioxide
NOx = oxides of nitrogen (NO, NO2)
NPR = nozzle pressure ratios
ODR = ordinary differential equation
OH = hydroxyl radical species
PDE = partial differential equation
PDF = probability density function
PFC = passive flow control
PID = proportional integral derivative
PIV = particle image velocimetry
POD = proper orthogonal decomposition
PRT = power resonance tube
PSE = parabolised stability equation
PVGJ = pulsed vortex generator jet
RANS = Reynolds-averaged Navier–Stokes
RF = radio frequency
RHS = right-hand side
rms, RMS = root mean squared
RQL = rich-quench-lean burn
SBLI = shock–boundary-layer interaction
SISO = single-input, single-output
SSL = suction surface length
SSTOL = super short take-off and landing
SVC = shock vector control
xxi

TAMI = tip air mass injection


T-S = Tollmien–Schlichting (instability)
TS = throat shifting
TVC = trapped-vortex combustor
TE = trailing edge
UAC = ultrasonically absorptive coating
UAV = uninhabited/unmanned aerial vehicle
UHC = unburnt hydrocarbon
URANS = unsteady Reynolds-averaged Navier–Stokes
VBB = vortex breakdown bubble
VG = vortex generator
VGJ = vortex generator jet
VSF = vortex shedding frequency
WC = wave cancellation
ZMF = zero mass flow
ZNMF = zero-net mass flux
Preface

Flow control technologies have been used for decades to control fluid flows, and
some employ underlying concepts that date back centuries. More recently, active
flow control terminology has been used with disparity among the different disci-
plines. Here, we use modern flow control to embrace the diverse active flow con-
trol technologies, terminologies, and disciplines. Modern flow control has become
an enabling technology as fluid dynamics, controllers, actuators, and sensors
merge to form advanced control systems capable of solving challenging aerospace
applications. The potential benefits from modern flow control have spawned major
research initiatives in government, industry, and academic sectors of aeronautics.
This text will present the current state of the art in modern flow control techno-
logies and highlight the application of these technologies to aerospace platforms.
The initial chapters serve to introduce the fundamentals of modern flow control,
including basic concepts, terminology, history, flow physics, actuators, sensors,
modeling/simulation, and instability and control theories. The later chapters cover
applications of flow control to current and next-generation air vehicle systems,
including fixed wing airfoils, turbomachinery, combustion, aeroacoustics, vehicle
propulsion integration, and rotorcraft. The text is not intended to be a thorough
review of each application; rather, it is focused on introducing the reader to the
various uses of modern flow control.

Ronald D. Joslin
Daniel N. Miller
May 2009

xxiii
Table of Contents

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

Acronyms and Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxiii

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxv

Chapter 1. Brief History of Flow Control . . . . . . . . . . . . . . . . . . . . . . . 1


David R. Williams, Illinois Institute of Technology, Chicago, Illinois; and
Douglas G. MacMynowski, California Institute of Technology, Pasadena,
California

Flow Control in the Empirical Era . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


Modern Flow Control: Leveraging Flow Instabilities . . . . . . . . . . . . . . . . . . . . . . . . 10
CFD Integration with Control Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Summary, Prospects, and Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Note on References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

Chapter 2. Physical Concepts Underlying the Development


and Application of Active Flow Control . . . . . . . . . . . . . . . . . . . . . . . 21
David Greenblatt, Technion–Israel Institute of Technology, Technion City, Haifa,
Israel; and Israel J. Wygnanski, University of Arizona, Tucson, Arizonia

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Fundamental Concepts in Historical Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Actuator and Actuation Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Free Shear Layer: A Prototype for Active Control . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Control of Jets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Controlling Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Wake Vortex Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

Chapter 3. Flow Control Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . 59


Douglas G. MacMynowski, California Institute of Technology, Pasadena, California;
and David R. Williams, Illinois Institute of Technology, Chicago, Illinois

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Control Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Flow Control Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

ix
x

Chapter 4. Role of Instability Theory in Flow Control . . . . . . . . . . . . 73


Vassilis Theofilis, Technical University of Madrid, Madrid, Spain

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Elements of Instability Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Multiple Role of Adjoints in Flow Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

Chapter 5. Dynamic and Closed-Loop Control . . . . . . . . . . . . . . . . . . 115


Clarence W. Rowley, Princeton University, Princeton, New Jersey; and
Belinda A. Batten, Oregon State University, Corvalis, Oregon

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
Classical Closed-Loop Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
State-Space Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
Model Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
Nonlinear Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

Chapter 6. Actuators and Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149


Louis Cattafesta and Mark Sheplak, University of Florida, Gainesville, Florida

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Actuators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

Chapter 7. Modeling and Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . 177


Christopher L. Rumsey and R. Charles Swanson, NASA Langley Research Center,
Hampton, Virginia

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Computational Fluid Dynamics Methodologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Computational Fluid Dynamics Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Exploration of Advanced Control Strategies Using Computational
Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

Chapter 8. Fixed Wing Airfoil Applications . . . . . . . . . . . . . . . . . . . . . 231


Avraham Seifert, Tel-Aviv University, Tel-Aviv, Israel; and Carl P. Tilmann,
Air Force Research Laboratory, Wright–Patterson Air Force Base,
Dayton, Ohio

Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
History and Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
Examples from the Present State of the Art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
xi

Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Summary and Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

Chapter 9. Turbomachinery Applications . . . . . . . . . . . . . . . . . . . . . . 259


Hermann F. Fasel and Andreas Gross, University of Arizona, Tucson, Arizona;
Jeffrey P. Bons, Ohio State University, Columbus, Ohio; and Richard B. Rivir
and Rolf Sondergaard, Air Force Research Laboratory, Wright–Patterson Air
Force Base, Dayton, Ohio

Introduction and Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259


Experimental Investigations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320

Chapter 10. Combustion Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321


Suresh Menon and Ben T. Zinn, Georgia Institute of Technology, Atlanta, Georgia

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Physics of Combustion Instability and Lean Blowout . . . . . . . . . . . . . . . . . . . . . . . . 324
Control of Combustion Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
Control of Lean Blow Out . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
Future Prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351

Chapter 11. Aeroacoustics of Flow Control . . . . . . . . . . . . . . . . . . . . . 353


William Devenport, Virginia Polytechnic Institute and State University, Blacksburg,
Virginia; and Stewart Glegg, Florida Atlantic University, Dania Beach, Florida

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
Sound Generation by Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
Leading Edge Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
Trailing Edge Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
Separated Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Acoustic Control of Jets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Circulation Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
Synthetic Jets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371

Chapter 12. Air-Breathing Propulsion Flowpath Applications . . . . . . 373


Daniel N. Miller, Lockheed Martin Aeronautics Company, Fort Worth, Texas;
and Jeffrey D. Flamm, NASA Langley Research Center, Hampton, Virginia

Tomorrow’s Propulsion Flowpath and the Need for Flow Control . . . . . . . . . . . . . . . 373
Applications to Tomorrow’s Propulsion Flowpath . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
Application to the Inlet Aperture: Shock/Boundary Layer Flow Control . . . . . . . . . . 378
Application to the Inlet Duct: Separation and Vortex Flow Control . . . . . . . . . . . . . . 383
Application to the Nozzle: Fluidic Thrust Vectoring Flow Control . . . . . . . . . . . . . . 391
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
xii

Chapter 13. Flow Control for Rotorcraft Applications . . . . . . . . . . . . 403


Ahmed A. Hassan, Sigma Technologies, Mesa, Arizona; Michael A. McVeigh,
Boeing Company, Philadelphia, Pennsylvania; and Israel Wygnanski,
University of Arizona, Tucson, Arizona

Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
Applications of Active Flow Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
Prospects and Challenges of Modern Flow Control . . . . . . . . . . . . . . . . . . . . . . . . . . 439

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513

Supporting Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523


Chapter 2

Physical Concepts Underlying the Development


and Application of Active Flow Control

David Greenblatt*
Technion–Israel Institute of Technology, Technion City, Haifa, Israel
and
Israel J. Wygnanski†
University of Arizona, Tucson, Arizonia

I. Introduction
Machines designed with the express purpose of diverting the natural path of a
flowing fluid, in order to achieve a secondary objective, have a history much longer
than we generally realize. In fact, the famous mechanically driven Archimedean
screw, used to raise water primarily for irrigation, dates back to the seventh century
B.C. (Dalley and Oleson, 2003). It is also believed that the first noria, used for
lifting water into an aqueduct and powered by the flow itself, may have been used
as early as the fourth century B.C. (Reynolds, 1983). These two ancient prototypes
furnished us with two inherently and conceptually different fluid machines,
namely the machine powered by an external agent (hand-operated Archimedean
screw) and a machine driven by the flow itself (flow-powered noria).
Over the intervening millennia, culminating in the early 20th century, mechani-
cally driven pumps and fluid-driven machines had become so commonplace that
they were, in fact, indispensable. In parallel, theoretical fluid mechanics succeeded
in providing explanations for a large number of observed phenomena, particularly
for irrotational flows, vortex motions and limited effects of viscosity (e.g., Lamb,
1932). Nevertheless, a schism remained between empirically based hydraulics
and mathematical fluid mechanics: the former was advancing the technology of
fluid machinery, whereas the latter was apparently incapable of explaining or
analyzing many hydraulics-based observations. The conceptual leap and resolution

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Senior Lecturer, Faculty of Mechanical Engineering. Senior Member AIAA.
†Professor, Department of Aerospace and Mechanical Engineering. Fellow AIAA.

21
22 D. GREENBLATT AND I. J. WYGNANSKI

was made by Prandtl, in his famous paper of 1904, where he not only defined the
physical and mathematical attributes of the boundary layer, but also invented so-
called boundary-layer control (BLC). The effect of BLC on aerodynamics was
particularly profound, and the degree to which this branch of science and engi-
neering progressed can be gauged from the landmark volumes edited by Lachmann
in 1961, which provided an exhaustive treatment of theoretical, experimental and
applied BLC methods. Unfortunately, BLC did not live up to its full potential, and
a variety of reasons were offered, from the overly heavy and complex plumbing
systems required for BLC, to the space race!
In recent decades, active BLC or, more generally, active flow control (AFC) has
re-emerged as a research area with potential application to a wide variety of prob-
lems of engineering interest. In contrast to traditional BLC (Lachmann, 1961),
present-day AFC relies on local active perturbations to bring about global changes
to the flowfield. Perturbations may be small relative to a characteristic velocity or
dimension, and thus exploit the unstable nature of the flow, or may be of the order
of the flow velocity or larger, and therefore “force” the flow. In both instances, the
objective is to produce flowfield changes that result in net improvements in perfor-
mance, reliability, safety, efficiency, etc. A central challenge for researchers and
engineers is how best to exploit hydrodynamic instabilities in order to attain a
meaningful benefit. During recent decades, AFC has evolved from empirical dem-
onstrations in canonical and relatively simple two-dimensional flows (e.g., Oster
et al., 1978) to applications-driven experiments and, in some cases, to full-scale
ground demonstrations (Kibens et al., 1999) and flight-testing (e.g., Musquere,
2003; Nagib et al., 2004; Phillips, 2003). Nevertheless, our understanding of these
flows is limited owing to their complexity, because turbulence coexists with so-
called coherent structures, which are usually driven by at least one instability
mechanism. Thus, theoretical methods (Gaster et al., 1985; Reau and Tumin, 2002)
and computational models (Rumsey et al., 2006) can at best describe only qualita-
tive trends and generally do not have a complete predictive capability. As a con-
sequence, the main advances have been empirical, or semi-empirical, and will likely
remain that way for some time. In some respects, we stand at a similar impasse as
that faced by hydraulicists and theoreticians at the turn of the 20th century.
The purpose of this chapter is to present a historical evolution of the funda-
mental concepts of modern AFC and to discuss some representative or proto-
typical applications. We generally limit our discussion to AFC that stems from the
introduction of hydrodynamic time-dependent (pulsating or periodic) perturba-
tions, although some spatially periodic perturbations are considered. Applications
discussed include control of mixing, jets, boundary layers (including improve-
ment of airfoil and wing performance and control of unsteady separation and
attachment and associated dynamic stall), wake vortices, three-dimensional flows,
and bluff bodies. This chapter does not address questions of feedback control (see
Chapters 3 and 5), and thus, from a strictly control-theory viewpoint, the discussions
here relate to open-loop, or feedforward, control.

II. Fundamental Concepts in Historical Perspective


Manipulation of the Navier–Stokes equations advanced the technology associ-
ated with continuum fluid dynamics and, via stability analysis, led to the concept
DEVELOPMENT AND APPLICATION OF AFC 23

of active flow control. However, fundamental discoveries that were crucial to our
understanding of turbulent flows inadvertently retarded the development of methods
to control them. One such discovery was Reynolds’ (1883) famous observation of
transition in pipe flows. He differentiated between quiescent (laminar) flow and
sinuous (turbulent) flow, and, because the latter seemed to be random, he also
separated the instantaneous velocity vector into steady and random components.
Time-averaging of the equations (Reynolds averaging) and the resulting Reynolds
stresses (Reynolds, 1894), led to a century of study aimed at developing viable
predictive techniques based on this viewpoint.
It should be remembered that turbulence represents the natural state of a flow in
most flows of engineering significance (e.g., Tennekes and Lumley, 1999): simply
consider the flow over any common vehicle, or within any pump, turbine or
engine. It is characterized by an irregular three-dimensional, vortical motion that
is accompanied by vigorous mixing of the fluid. Mixing is an inseparable ingredient
of turbulent flow, because an irregular motion by itself can also occur in solids,
particularly compliant ones. The irregularity of the motion implies the existence
of a wide spectrum of scales, suggesting that a complete deterministic description
of the flow is not attainable. It is hardly surprising, therefore, that turbulence was
described in statistical terms that decomposed the velocity and the pressure into
mean and fluctuating components. This decomposition created a new set of equa-
tions that resemble the instantaneous equations of motion, but which unfortunately
cannot be solved unless additional equations for new unknowns created by this
decomposition (i.e., the Reynolds stresses) are somehow determined. For many
decades, this was the heart of the “turbulence problem” or, as it was often referred
to in the literature, the “closure problem.” The generation of additional equations
from the original Navier–Stokes equations never resolved this dilemma, because
it always resulted in a larger number of unknowns than equations being available.
It simply shifted the inevitable ad hoc decision to other terms. An entire “modeling
industry” that concentrated on providing mathematical models of turbulence
evolved over the years. This approach has effectively established Reynolds-
averaged Navier–Stokes (RANS) methods as the main practical tool available to
date, and it is widely used in industrial applications. The success of these models
depends to a large extent on the quality of the empirical input that they use. They
are therefore capable of “postdicting” the type of flows about which there is a
substantial amount of information, rather than “predicting” the behavior of an
entirely novel flow.
Since this approach does not explain the physical processes governing turbulent
shear flows, nor does it indicate the means of manipulating or controlling such
flows, traditional Reynolds averaging is considered to be detrimental to the control
of turbulent shear flows. Thus, the most significant discovery of Osborne Reynolds
led to a fatalistic approach to the control of turbulence.
In the context of AFC, numerical simulation of the entire flowfield by applying
finite difference or spectral methods to the instantaneous equations of motion has
the potential of becoming a major utility. However, numerical simulation, like an
experimental facility, does not delineate the parameters affecting the flow according
to their relative significance. It therefore provides results in an indiscriminate
fashion, with no insight into the physical aspects governing the flow. However, if
the physical parameters affecting the flow are even vaguely understood, then the
24 D. GREENBLATT AND I. J. WYGNANSKI

results provided by direct numerical simulation (DNS) may become a very valuable
asset. Indeed, the results of DNS have become a major source of turbulence data
that cannot be obtained from experiments.
As noted in Chapter 1, the most important advance leading to the AFC concept
was Prandtl’s boundary-layer theory. It separated the flow field into a thin layer of
rotational fluid adjacent to a solid surface, surrounded by a large body of irrota-
tional flow that can be considered to be inviscid. The concept describes the flow
around streamlined bodies very well, and it also explains the frictional losses and
the convective heat transfer occurring between the surface and the adjacent fluid.
The major practical simplification stemming from boundary-layer theory is the
ability to predict the pressure distribution around a streamlined body using irrota-
tional flow solutions. This also led to the development of the separation concept,
where the fluid retarded by viscous forces breaks away from the surface as a result
of a strong adverse pressure gradient. Experiments supported Prandtl’s assump-
tions and provided criteria for the breakdown of these assumptions due to separation.
They also revealed the conditions under which steady, laminar boundary layers
develop instabilities and become turbulent; the concept of instability is covered in
detail in Chapter 4. Prandtl’s concept provided the framework for the computa-
tional viscous/inviscid interaction concept by linking far-field potential-flow
solutions with viscous ones at the edge of the boundary layer. Criteria for the
stability of the boundary layer and its existence, exemplified by the criteria for
separation, followed naturally from this.
Initial extensions of boundary-layer theory to turbulent flow using turbulence
models based on Reynolds decomposition hindered progress in AFC, because it
suggested that the random motion was determined by local flow conditions and
therefore control at a specific location would not carry long-term effects with it.
This was also tied to the modeling that was extensively used in engineering appli-
cations and that led to the famous statement that “turbulent flow forgets its origin.”
This belief was deeply engrained until the discovery of large coherent structures
in turbulent shear flows.
Although the Reynolds-averaged equations are mathematically correct, their
applicability is limited to those regions in turbulent shear flows where the turbu-
lence may be assumed to be approximately homogeneous and isotropic. In an
intermittently turbulent flow such as exists in the outer part of a turbulent boundary
layer, a wake, or a jet, the Reynolds-averaged equations will lump together and
indiscriminately average the vortical (turbulent) fluctuations with the irrotational
fluctuations existing outside the instantaneous turbulent boundaries of the flow.
Corrsin and Kistler (1955) recognized this shortcoming and introduced the
concept of a “superlayer,” which represents a thin, highly contorted boundary
separating the turbulent from the irrotational zones. Vorticity is imparted to the
irrotational fluid along this boundary through the action of viscosity. By assuming
that the superlayer is continuous without islands of turbulent fluid being present
in the irrotational zone, Corrsin and Kistler were able to measure the duration T of
the large eddies at the outer edge of the boundary layer, determining that
TU• /d  2.5. This was probably the first measurement delineating the average
size of a large eddy propagating at the outer edge of the boundary layer and indi-
cating that turbulence is not as random as it was previously believed to be.
Kovasznay et al. (1970) introduced the notion of zone averaging and conditional
DEVELOPMENT AND APPLICATION OF AFC 25

sampling, which exposed experimentally the limitations of a purely statistical


approach to turbulence. Kline et al. (1967) revealed the inhomegeneity existing in
the wall region of a turbulent boundary layer, and Blackwelder and Kaplan (1972)
demonstrated the coherence of some large-amplitude fluctuations across the entire
boundary layer. The most significant observations that gave the impetus to research
on coherent structures in turbulence were made by Brown and Roshko (1971,
1974) in a two-dimensional mixing layer. Their schlieren photographs showed
that the turbulent mixing layer is dominated by large-scale eddies, which transport
within them much smaller and approximately homogeneous turbulent eddies. The
mixing-layer spreading rate could clearly be related to the growth of the large
eddies, which engulf irrotational fluid from the surrounding streams.
Research on coherent structures in turbulence switched into high gear and
dominated experimental and theoretical investigations during the 1970s and
1980s. Zone averaging and conditional sampling soon gave way to more sophisti-
cated variable-interval time-averaging (VITA) techniques in which some temporal
information relative to a well-recognized event was maintained and used.
Experimental methods were altered, with flow visualization and particle image
velocimetry (PIV) techniques coming to the foreground because they provide
instantaneous information over a large region in the flow rather than detailed
temporal information concentrated at a single point. Flow visualization proved to
be very helpful in formulating new ideas, and quantitative measurements enabled
the researcher to check these ideas. One may realistically expect the technology
for an instantaneous and complete mapping of the velocity field become routinely
available in the near future.
Extending Reynolds’ original idea to the so-called triple decomposition
(Hussain and Reynolds, 1970; Reynolds and Hussain, 1972) provides a formalism
for tackling the problem of turbulent shear-flow control. The triple decomposition
recognizes that the unsteady motion may be decomposed into large, coherent,
deterministic structures that are predictable and smaller ones that presently cannot
be predicted and are described by statistical methods and therefore presumed to be
random. This approach may provide the theoretical tools necessary for controlling
the large eddies, and hydrodynamic stability theory, applied to a turbulent flow
field, may provide the first step in developing rational models for flow control.
Indeed, when inviscid stability theory was applied to the forced turbulent mixing
layer (Gaster et al., 1985) it provided good predictions of the relative amplitude
and phase distribution across the mixing layer, but rendered only a qualitative
prediction of the perturbation amplification in the direction of streaming. One
may thus view large coherent structures as a product of interacting instability
waves that propagate downstream while either amplifying or decaying during the
period of time under consideration, although this view seems to apply best to
flows that are inviscidly unstable. Nevertheless, so-called random turbulence also
plays a role in determining the quantitative development indicators. Here we have
no alternative but to resort to a model of some kind.
The ideas of control and manipulation of turbulent flow originated from the use
of a low-level forcing signal that was needed to provide a phase reference for data
acquisition that employed a limited number of hot-wire probes, thus giving infor-
mation at only a select number of locations. It was soon realized that even a low-
amplitude disturbance alters the mean flow and the intensity and distribution of
26 D. GREENBLATT AND I. J. WYGNANSKI

the turbulent fluctuations (Oster et al., 1978). This led the way to the control of
other free shear flows, the control of separation, and the isolation of some basic
controlling parameters. In addition, small-scale turbulence proved to be very
sensitive to the modes of interaction of the coherent structures, and thus control
over instability modes also enables the control of chemical reaction rates, which
implies control at the molecular level (Roberts, 1985).
Recent decades have witnessed an unprecedented growth in AFC studies and
projected applications. When our ability to control these flows matures, the
resulting changes to perceived characteristics of turbulent shear flows will
undoubtedly have a major technological impact, as it will alter the dimensions
and shapes of wings, diffusers, combustion chambers, ground and underwater
vehicles, etc. In short, any machinery associated with fluid flow, either external
or internal, is bound to look different in the future. Forty years have passed since
flow visualization exposed the existence of large coherent structures in turbulent
shear flows. During these years, the direction of turbulence research has changed
dramatically from statistical compilation of turbulence intensities and modeling
of Reynolds stresses to a search for coherent structures, and from accepting the
inevitability of the existence of a “universal” shear flow to manipulation and
alteration of that “universality.” The time has come to generate some consensus
on the subject and present a point of view, which, it is to be hoped, will guide
the next generation of graduate students and will slowly filter through the indus-
trial establishment.

III. Actuator and Actuation Basics


Flow excitation, actuation, or forcing is a critical aspect in transitioning active
control from the laboratory to real-world applications. On the one hand, actua-
tors with sufficient authority must be developed that simultaneously provide a
net system benefit. On the other hand, the correct location, frequency, orienta-
tion, type of actuator, etc, must be determined, and here theoretical studies are
only partially helpful. The earliest demonstrations of excitation-based control in
the laboratory relied on the use of loudspeakers to generate acoustic waves
(Ahuja et al., 1983; Freymuth, 1966), as they are inexpensive, convenient, and
robust, with simple and precise control of frequency, relative phase, and ampli-
tude. Nevertheless, their lack of control authority per unit weight or volume
limited their potential for real-world applications. The search for alternate actua-
tion methods has led to the development of many new methods specifically aimed
at AFC. Common examples are zero or nonzero (pulsed) mass-flux devices,
including voice-coil-based (e.g., McCormick et al., 2001; Nagib et al., 2007),
piezoelectric-driven (Chen et al., 2000), and valve- or siren-type (Bachar, 2001;
Seifert and Pack, 1999; Seifert et al., 1996) devices; rigid benders (Neuburger
and Wygnanski, 1987); piezoelectric flappers (Seifert et al., 1998); rotating wall-
elements (Viets et al., 1987); oscillating wires (Bar-Sever, 1989); mechanical
surface oscillators (Park et al., 2001); combustion-driven devices (Crittenden
et al., 2001); electrohydrodynamic devices, such as pulsed corona wires (Sosa
et al., 2006), dielectric barrier discharge electrodes (Enloe et al., 2006; Post and
Corke, 2004, 2006), and arc filament actuators (Samimy et al., 2004a, b); and
DEVELOPMENT AND APPLICATION OF AFC 27

magnetohydrodynamic Lorentz-force actuators (Cierpka et al., 2007; Weier and


Gerbeth, 2004).
With this vast and ever-growing range of actuation methods, a key task is to
establish a common “output” parameter so that their relative effects on the flow or
performance parameters of interest can be compared. The parameter most widely
used to characterize the output of an actuator is the so-called momentum
coefficient:

J ·J Ò
Cm ,tot = Cm + · Cm Ò = + (1)
q• L q• L

where J and ·J Ò are the steady and unsteady actuator momentum addition,
respectively (e.g., Greenblatt and Wygnanski, 2000; Seifert et al., 1996). In most
cases, the momentum components cannot be predicted from first principles—
although lumped-element or reduced-order modeling is used (e.g., Gallas et al.,
2002; Sharma, 2007)—and the actuator must be calibrated by directly measuring
the velocity field and hence its momentum components. A notable exception is
Lorentz-force-type actuators, where the body force momentum is directly gener-
ated in the fluid and calibration can be conducted from first principles (e.g., Weier
and Gerbeth, 2004). A more detailed discussion of actuators is provided in
Chapter 6.
The parameter ·CmÒ is not without its limitations. For zero-mass-flux devices,
the suction phase does not directly add momentum, but removes low momentum
fluid from the near-wall region and reconfigures the local vorticity. It is also gen-
erally assumed that calibrations performed under quiescent conditions (U• = 0)
are valid under test conditions (U• π 0) (see the discussion in Greenblatt et al.,
2006b). The preferred method of pulsed control is to superimpose a net positive or
negative steady mass flux onto a nominally zero-mass-flux device; certain pulsed
valves have the disadvantage that the relative proportion of momentum cannot be
varied. Some actuators operate at or near resonance, corresponding to frequencies
that are very much higher than those required for exciting a useful instability. In
these instances, some form of low-frequency modulation is employed, and ·CmÒ is
not the only parameter characterizing the actuator input. The validity of ·CmÒ in the
context of large-amplitude forcing, where uj > U•, has also been questioned by
Kiedaisch et al. (2006).

IV. Free Shear Layer: A Prototype for Active Control


Control of the two-dimensional incompressible turbulent mixing (free shear)
layer, provides perhaps the simplest AFC paradigm and is a natural starting point
for illustrating basic principles and defining control parameters. In the absence of
perturbations, the layer spreads out in the direction of streaming, because the large
spanwise structures that it contains engulf fluid from both streams, rolling it into
a discrete array of vortices. Most of the orderly, traveling structures represent
concentrations of spanwise vorticity that are associated with the Kelvin–Helmholtz
(K–H) instability mechanism discussed below and in Chapter 4 (e.g., Cohen and
28 D. GREENBLATT AND I. J. WYGNANSKI

Wygnanski, 1987a; Gaster et al., 1985). Naturally generated K–H eddies are not
uniform in size, possessing a range of wave numbers. Consequently, the mean
flow, which gives rise to their development and is commensurate with their size,
spreads out linearly in the direction of streaming. The introduction of a small
periodic mechanical excitation at the flow origin (e.g., Oster et al., 1978) accelerates
and regulates the generation of large coherent structures, transferring high-
momentum fluid across the mixing layer. From the perspective of AFC, we are
exploiting the K–H instability mechanism to control the spreading rate and turbulent
mixing (e.g., Cohen, 1985; Oster et al., 1978). The sole difference between the
two photographs shown in Figs. 1a and 1b, illustrating a perturbed mixing layer
that is visualized by smoke, is that the frequency of the excitation in Fig. 1b is
twice that shown in Fig. 1a. The eddy size and the corresponding width of the
mean flow are clearly illustrated by the smoke visualization.

A. Two-Dimensional Perturbations
The spreading rate of a mixing layer that is excited at a frequency fe by a plane wave
emanating from the trailing edge of a solid partition or nozzle can be divided into three
regions, two of which are visible in Fig. 2a, which represents the results in dimensional
form. Region I, which is adjacent to the trailing edge of the partition (Fig. 2b), diverges,
principally as a result of the amplification of quasi-two-dimensional waves. Region

II, corresponding to q + ∫ q fe /U ª 0.075, starts where the mixing layer ceases to

Fig. 1 Periodically excited, plane turbulent mixing layer generated between two
parallel streams. The upper streams (a) are excited at half the frequency of the lower
streams (b) (Oster et al., 1978).
DEVELOPMENT AND APPLICATION OF AFC 29

a)

LIVE GRAPH
Click here to view

b) 0.2

III
0.15
q+

0.1
II
I

0.05

LIVE GRAPH
Click here to view Data Scatter
0
0 1 2 3 4 5
Rf+

Fig. 2 Spreading rate of a mixing layer subjected to periodic excitation: a) spreading


rate as a function of amplitude for a fixed frequency (from Oster and Wygnanski,
1982), and b) spreading rate in dimensionless form (adapted from Wygnanski and
Petersen, 1987).

grow because it is neutrally stable to the imposed harmonic disturbances that


dominate the flow. q is the momentum thickness, defined as

U - U2
• Ê U - U2 ˆ
q= Ú
-• U - U
1 2
ÁË1 - U - U ˜¯ dy
1 2
(2)
30 D. GREENBLATT AND I. J. WYGNANSKI


fe is the excitation frequency, and U = _12 (U1 + U2), where U1 and U2 are the velocities
of the two streams. Disturbances of longer wavelength amplify over a greater
distance and therefore dominate the flow further downstream. The physical down-
stream location corresponding to the onset of neutral stability is thus inversely
proportional to the frequency of the excitation. Wherever a single frequency domi-
nates the flow, region II extends over the approximate range 1 < Rf + < 2, where

f + = fx/U = x/lx represents a dimensionless distance measured in terms of the
number of streamwise wavelengths that it contains and R = (U1 - U2)/(U1 + U2)
represents the strength of the spanwise eddies (Fig. 2b). Beyond Rf +  2, the
coherent Reynolds stresses (associated with the excitation frequency) lose their
relative significance and the mixing layer continues to spread out linearly (region
III) with increasing distance from the splitter plate.
The spreading rate, and hence the entrainment, depends strongly on both fre-
quency and amplitude. A dimensionless frequency can be clearly defined, but the
correct scaling for the perturbation amplitude remains an open question. The
utility and limitations of purely theoretical approaches to shear-layer predictions
can be summed up by comparing linear inviscid stability theory, incorporating a
first-order correction for slow spatial variations of the mean flow, with experimental
data. Here, the comparison was based on the amplitude of the integral of the
modulus of the fundamental component of the artificially excited wave train at
each streamwise station (Fig. 3).
This comparison of the overall amplification of the perturbation indicates that
inviscid theory is only capable of predicting the growth trend. However, when the
comparison is made on a purely local basis, agreement in both the amplitude and

10

R = 0.25, L0 = 48 mm
∫–•ÍucÍdy

1

R = 0.43, L0 = 38 mm
Normalized

R = 0.43, L0 = 76 mm

1
0 1
x/L0

Fig. 3 Overall amplification of disturbances with distance from the trailing edge
(St ª 0.22). Symbols represent experimental data, the dashed line represents linear
inviscid stability theory, and the solid lines represent linear inviscid stability theory
incorporating nonparallel effects (Gaster et al., 1985).
DEVELOPMENT AND APPLICATION OF AFC 31

2.5
x/L0=0.249 0.349 0.448 0.548 0.698 0.847
y/L0

R = 0.25, L0 = 48 mm
–2.5
Normalized uc

Fig. 4 Lateral distribution of the coherent amplitude variation for the forced free
shear layer at various x/L0 stations, St ª 0.22 (Gaster et al., 1985).

phase distribution across the mixing layer is good (e.g., Fig. 4). A principal limitation
of the inviscid linear model is that it fails to account for what is usually referred to as
“fine-scale” turbulence. In addition, there is no valid basis for neglecting nonlinear
terms, because approximately 45% of the energy associated with the streamwise
fluctuations occurs at the forcing frequency and therefore these levels of velocity
fluctuation cannot be considered as small perturbations. The good predictions of
phase-locked velocity fluctuations (e.g. Fig. 4) apparently result from the fact that the
eigenfunctions of the fundamental part of the motion are given by a linear approxi-
mation in the solutions of weakly nonlinear stability problems. It is only through
the balance of terms that an amplitude scaling is defined. This may explain the
relatively poor amplification predictions, which are limited by the additional factors
of dissipation and energy transfer to fine-scale turbulence (Hussain, 1983).

B. Three-Dimensional Development
If the remaining incoherent (turbulent) motion is not only random but also fine-
scale, it may be represented by a simple eddy viscosity model (Reau and Tumin,
2002; Marasli et al., 1992) that allows prediction of changes to the spreading rate
and the associated coherent Reynolds stresses resulting from the excitation. However,
this model does not consider the possible coherence of the streamwise streaks that
ride on the spanwise structures and are stretched by the strain field existing between
adjacent rolls. Shadowgraph pictures by Konrad (1976) revealed the existence of
these streaks, whose spanwise spacing increased with increasing distance from the
splitter plate. This spacing seems to scale with the local width of the flow, suggesting
that the streaks are a product of a genuine instability of the mixing layer and their
contribution to the enhancement of mixing may be significant.
Three-dimensional motions in a mixing layer were first studied by Bernal and
Roshko (1986) through flow visualization in water. They suggested that the secondary
32 D. GREENBLATT AND I. J. WYGNANSKI

streamwise vortices are simply portions of a warped vortex that threads its way up
and down between adjacent spanwise vortices, thus changing its streamwise direc-
tional sign on each pass. Further observations of streamwise vortices in a forced,
plane mixing layer were carried out by Huang and Ho (1990), Tung and Kleis (1996),
Leboeuf and Mehta (1996), and others.
Fiedler et al. (1998) delineated three aspects or categories of three-dimensionality:
natural secondary three-dimensional development of the two-dimensional
shear layer (discussed above); basic flow three-dimensionality; and geometric
three-dimensionality. Thus, from an AFC perspective, we can conceive of 1)
three-dimensional perturbations of two-dimensional flows; 2) two-dimensional
perturbations of inherently three-dimensional flows; and 3) three-dimensional
perturbations of three-dimensional flows, which is clearly the most general and
complex case. Some aspects of item 1 are discussed next.

C. Three-Dimensional Perturbations
Nygaard and Glezer (1991, 1994) were able to excite the spanwise rolls and the
streamwise streaks in a time-dependent manner in water, by embedding heating
elements in the splitter plate separating the two streams. This enabled independent
control of the spanwise and streamwise wavelengths at low Reynolds number.
Flow visualization suggested that the streamwise streaks had a “L” shape that
appeared first near the high-speed edge of the primary spanwise wave before it
rolled up into a vortex. Direct numerical simulations of a temporally growing
mixing layer were first performed by Metcalfe et al. (1987), who observed that
spanwise instability modes lead to the formation of pairs of counter-rotating
streamwise vortices in the braids connecting adjacent spanwise rolls.
Interest in the interaction between spanwise and streamwise vortices in aero-
dynamics has increased recently, particularly as a result of the relative success of
lobed and chevron nozzles employed for jet noise reduction. There are other
applications where spanwise and streamwise vortices interact, and such interaction
may dominate the mean flow. For example, in a wall jet flowing over a convex
surface, there is an interaction between the K–H spanwise rolls and streamwise
vortices that are generated by a centrifugal instability. In this case at least, the
origin of the streamwise vortices is known but not the ensuing interaction with the
spanwise rolls.
Lasheras and Choi (1988) enhanced and regulated the streamwise structures by
corrugating the partition separating the two streams or by indenting its trailing
edge in a sinusoidal manner. This procedure enabled them to control the initial
spanwise wavelength of the streaks and to observe their intensification in the
direction of streaming.
Kit et al. (2007) attached an oscillatory, chevron-shaped fliperon to a stationary
partition in a mixing layer having a velocity ratio of 0.4 (R = 0.429). It generated
spanwise rolls that were initially parallel to the chevron trailing edge, creating a
spanwise periodic bending and swelling of the large eddies. The bending of the
large eddies was achieved by self-induction, which created pairs of counter-rotating
streamwise vortices whose relative intensity was mostly affected by the shape and
size of the chevron fliperon and not by its oscillations. The flow at the trailing
edge of the splitter plate was fully turbulent, having a Reynolds number based on
DEVELOPMENT AND APPLICATION OF AFC 33

initial momentum thickness exceeding 90. However the initial momentum thick-
ness, qI, did not properly represent the width of the flow whenever the mixing
layer’s velocity profile contained a wake component generated by the merging
boundary layers of the splitter plate.
The center of a plane mixing layer is traditionally defined by the lateral loca-

tion y0, where the mean velocity is U = 1_2 (U1 + U2). This location may also repre-
sent the lateral locations of the centers of the large spanwise rolls and enables the
assessment of the degree of their distortion. The distribution of y0 for the station-
ary chevron fliperon is presented in Fig. 5a and compared with the plane mixing-
layer data obtained in the same facility under identical flow conditions. The
two-dimensional results are indeed independent of the spanwise coordinate, and
are represented by a single straight line in Fig. 5a. The initial three-dimensional
perturbations generated by the stationary chevron fliperon resulted in the most
significant bending of the spanwise rolls, starting some 300 mm downstream of
the trailing edge. The bend increased monotonically until x = 800 mm, where the
difference between the location of y0 opposite the notch and the cusp in the
fliperon was approximately 25 mm. This distance is commensurate with the depth
of the notch, but more surprising is the perseverance of the distortion in the direc-
tion of streaming.
A representative local width of the flow q(z) is shown in Fig. 5b for the two-
dimensional and the chevron trailing edges at two spanwise locations: one opposite
the cusp in the fliperon and the other opposite its notch. The two-dimensional fli-
peron in the absence of excitation generates a traditional two-dimensional mixing
layer, whose spreading rate dq/dx is practically constant over the entire range of
measurement. The stationary chevron fliperon created a mixing layer whose width
q varied along the span but was everywhere larger than the momentum thickness
of the corresponding plane mixing layer. The initial rate of divergence dq/dx
opposite the notch in the chevron was 50% higher than for the corresponding
two-dimensional trailing edge. The implied increase in mixing by the introduction

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view
a) 0
2-D, 3-D fliperon
400 800 x (mm) b) 0.12
0 400 800 x (mm)
0.24
3-D A = 0, fliperon notch 2-D, 3-D fliperon
0.20 3-D A = 0, fliperon cusp 0.10

0.16
0.08
–y0/λx
0.12 θ/λx 2-D, A = 0,
2-D A = 0,
0.06 linear fit
linear fit
0.08
3-D A = 0, fliperon notch
0.04
0.04 3-D A = 0, fliperon cusp

0 0.02

–0.04 0
0 1 2 0 1 2
Rx/λx Rx/λx

Fig. 5 Variation of y0 and q with distance from the trailing edge of a two-dimensional
fliperon and a chevron-shaped fliperon (Kit et al., 2007).
34 D. GREENBLATT AND I. J. WYGNANSKI

of streamwise vortices through tailoring the shape of the trailing edge is achieved
by spatially periodic, but static, perturbations. The resulting control is fundamentally
different to the enhancement of spanwise eddies by periodic excitation, and serves
as an indication of why chevron nozzles have such a large impact on the turbulent
structure of the jet and its associated noise.

V. Control of Jets
Interest in the control of jets by means of periodic perturbations has been driven
by the desire to increase mixing, reduce jet engine noise, and “vector” the jet
thrust. Enhanced mixing is exploited to control heat transfer, chemical reaction
rates, and jet plumes (e.g., Rice and Zaman, 1987). For jet noise control, both
subsonic and supersonic cases are relevant: for subsonic jets, instability waves do
not directly radiate sound, but they do drive the formation of sound-generating
turbulence (Moore, 1977); in supersonic jets, it is the instability waves that them-
selves radiate significant sound (e.g., McLaughlin et al., 1975). In addition, the
control of subsonic jets is also believed to have direct consequences for supersonic
jets, as the relationship between phase velocity and excitation Strouhal frequency
fe D/U0 appear to follow the same general trends observed in subsonic jets
(Lepicovsky et al., 1985).
Control of jets by periodic perturbations introduced at the lip of the jet has
followed two main approaches. The first involves the introduction into a base flow
of low-amplitude perturbations that are consistent with linear amplification cri-
teria; the second involves intermediate- or large-amplitude forcing that can be of
the order of the base flow itself, and at frequencies that are not necessarily related
to the base-flow stability characteristics (e.g., Parekh et al., 1996). The perturba-
tion can also be superimposed on the mean, or core, flow: when perturbations are
large, these are often called pulsed jets, a special case of which is the zero-mass-flux
jet, which is also termed a “synthetic jet” (Glezer and Amitay, 2002).

A. Small-Amplitude Perturbations
Low-amplitude axisymmetric perturbations produce an effect much like that on
the simpler two-dimensional shear layer, where traveling instability waves grow
in the directions of streaming until they reach a finite amplitude and roll up into
vortices. A jet can also be excited by a second instability wave, for example its
subharmonic (Kelly, 1967; Paschereit et al., 1995). Unlike the simple shear layer,
the axisymmetric shear layer is also unstable to azimuthal (or helical) modes. For
example, the jet is unique in the sense that both axisymmetric (m = 0) and helical
(e.g., m = ±1) modes can be excited simultaneously, thus exerting substantial
authority over the entire jet development.
Long and Petersen (1992) demonstrated that the shape of a jet could be dramati-
cally altered by the introduction of two opposite-signed helical perturbations. The
shape and orientation of the jet distortion could be predicted based on the
standing-wave pattern. Square and elliptical jets were produced in this way and
the spatial distribution of the coherent large-scale motion was documented. An
example is shown in Fig. 6, which illustrates the mean-flow distortion achieved by
introducing two spinning waves with mode numbers m = 1 and m = –1 into an
DEVELOPMENT AND APPLICATION OF AFC 35

Fig. 6 Effect of phase offset on jet cross section at x/D = 4.0: mean velocity contours
from 0.3Uj to 0.9Uj in 0.2Uj intervals. The relative phase offset between waves with
spinning mode numbers m = +1, -1 is a) 0 deg, b) 90 deg, c) 180 deg, and d) 270 deg
(Long and Petersen, 1992).

axisymmetric jet. This jet is comparable to that issuing from a 2:1 elliptical
nozzle. It can be seen from Fig. 6 that the primary axis of the elliptical jet is fully
controllable by merely varying the phase difference between the two helical
perturbations.
Challenges associated with high-speed jet control are formidable owing to the
difficulty in procuring actuators that can produce perturbations with sufficient
amplitude and bandwidth. For jet engine applications, these must also be small,
light, and rugged enough to withstand the harsh jet exhaust. Samimy et al.
(2004a–c) are developing a class of actuators termed localized arc filament
plasma actuators that appear to fulfill these stringent requirements, and have
been tested on a subsonic jet at Ma = 0.9 and on an ideally expanded jet at
Ma = 1.3. Preliminary results, which include limited flow images and far-field
acoustic data, suggest that the large-scale structures could be forced at fD/U0 < 1
and fD/U0 > 3.

B. Intermediate- and Large-Amplitude Perturbations


In many industrial applications, primary jets or boundary layers are subjected to
crossflowing or coflowing secondary jets, whose amplitude is not small in relative
terms. When these secondary jets are pulsed or zero-mass-flux, dramatic changes
36 D. GREENBLATT AND I. J. WYGNANSKI

to the primary or secondary flow can result. However, these perturbations do not
grow in the direction of streaming, and hence we refer to them as intermediate- or
large-amplitude perturbations. In the former case, perturbations are typically less
than 50% of the core flow, and in the latter case, they are of the same order as the
core flow or larger. These secondary jets or actuator perturbations are not necessar-
ily related to the linear stability characteristics of the flow.
Experiments by Reynolds and co-workers, commenced in the 1980s and sum-
marized in Reynolds et al. (2003), demonstrated that dramatic authority could be
exerted over a circular jet by combining an axial perturbation at a frequency fa and
amplitude about 0.2U0 with a helical perturbation at a lower frequency fh and
amplitude about 0.04D (i.e., dual-mode excitation). Helical perturbation (namely,
wobbling the jet by driving it in eccentrically and “weak acoustic forcing”)
resulted in successive vortex rings that were slightly displaced radially, producing
successive rings that were eccentric with respect to one another. Reynolds et al.
(2003) further showed that if fa /fh = 2, then the vortex rings will be tilted alter-
nately to opposite sides and, with sufficient large perturbations, the jet will bifurcate
(i.e., split into two) and appear to flap. The explanation is that the eccentric rings
tend to tilt one another by mutual induction, causing both to move away from the
common axis, stretching the jet core fluid between them. When the ratio fa /fh is an
integer, the jet can be made to divide into several separate streams. However, if the
ratio is not an integer, then no vortex ring exactly follows the one previously shed,
and the jet can be made to “bloom” in an amazing shower of vortex rings.
A dramatic example of this is shown in Fig. 7 for the conditions faD/U0 = 0.5 and
fa/fh = 2.4. For intermediate-level perturbations, the technique appears to be lim-
ited to low-Reynolds-number (<104) flows, because the thinner shear layer pro-
duces higher-frequency, more closely spaced vortices, which inhibits the mutual
induction mechanism discussed here.
In a study by Parekh et al. (1996), large-amplitude zero-mass-flux jets were
used to force the primary jet at frequencies that were one order of magnitude

Fig. 7 Blooming water jet at Re ª 20,000 (Reynolds, 2003).


DEVELOPMENT AND APPLICATION OF AFC 37

larger than the most-amplified frequency. Amplitude modulation was also used to
excite the jet at frequencies below the resonant frequency of the actuator. Wiltse
and Glezer (1993) extended this work to study supersonic jets, and pulsed jets
have also been employed to demonstrate effective mixing on a full-scale jet engine
(Kibens et al., 1999). One drawback of producing large-amplitude perturbations
for AFC is the associated heavy plumbing often used for the excitation jets.
In some instances, the secondary-jet amplitude considerably exceeds the core
flow, as in the case of a subsonic jet discharging perpendicularly into a boundary
layer. With an average jet-to-crossflow velocity ratio R = 2.58, M’Closkey et al.
(2002) showed that the jet penetration and downstream development and mixing
were strongly dependent on frequency, waveform, and duty cycle. Subsonic jet
vectoring was studied by Pack and Seifert (2001a, b), using large-amplitude zero-
mass-flux perturbations to attach flow to a diffuser attached to the jet exit. Jet
vectoring up to 10 deg was achieved by forcing a fraction of the jet circumference,
thereby locally enhancing the spreading rate and tendency of the jet to attach
locally to the diffuser in the vicinity of forcing. Jet vectoring was also achieved by
using a coflowing zero-mass-flux jet (Smith and Glezer, 2002), without employing
a diffuser or collar. Vectoring was achieved, as the primary jet was directed into
the zero-mass-flux jet orifice and the interaction between the jets produced a
closed recirculating flow domain.

VI. Controlling Separation


Control of separation is vitally important to the design of air vehicles (and indeed
all fluid machinery) and, together with transition, is a major factor limiting modern
designs. A central question regarding separation control relates to the optimum
frequencies and spanwise nodes for which a prescribed objective can be obtained.
Moreover, a distinction must be made between different objectives. For example,
are we trying to force attachment of a separated flow or prevent separation of a flow
that is already attached? Are we attempting to maximize CL,max to reduce landing
speed, or attempting to maximize L/D during cruise, or attempting to increase post-
stall lift and unsteadiness? In the following sections, we deal with some questions
relating to the generation of high lift, the effects of streamwise curvature, electro-
and magnetohydrodynamic methods, three-dimensional effects, unsteady separa-
tion and attachment, and the effects of Reynolds and Mach numbers.

A. High Lift Generated by Simple Devices


The attachment of an initially separated shear layer to an inclined solid surface
by means of small-amplitude perturbations is directly related to the shear layer’s
receptivity to the perturbations and their amplitude (Greenblatt and Wygnanski,
2000). Nishri and Wygnanski (1998) established that, for a fully turbulent upstream
boundary layer separating from a generic flap (i.e., an inclined surface), and in the
absence of surface curvature, the optimum reduced frequency (i.e., that requiring
the smallest ·Cµ Ò) was F +  1.2, where F +  feXte /U•. However, the optimum
F + required to prevent the separation of an initially attached flow was somewhat
higher, in the range 2 £ F + £ 4. It has long been recognized that the introduction
of a two-dimensional perturbation close to the separation line on airfoils can
38 D. GREENBLATT AND I. J. WYGNANSKI

produce substantial changes to post-stall lift, dramatic reductions in post-stall


unsteadiness, and improvements in CL,max and L/D (Greenblatt and Wygnanski,
2000). However, there is very little consensus on a specific optimum F+, with the
vast majority of investigations citing effective control at around 0.3 £ F+ £ 4 for
conventional low Reynolds number, providing that an amplitude corresponding to
·Cµ Ò  0.1% is exceeded. It is generally accepted that the maximum lift attainable
on airfoils and wings will depend on preventing separation from an initially
attached boundary layer, although the airfoil presents some additional challenges
when compared with the idealized deflected surface. For example, if control is
applied at or near the leading edge, then, in general, the flow will be laminar or
transitional and may be subjected to relatively large streamwise curvature. If
control is applied at a flap shoulder, then the upstream turbulent boundary layer is
subjected to an adverse pressure gradient that intensifies with the angle of attack.
An important, and particularly challenging, objective of separation control is to
provide a viable simplified alternative to present-day high-lift systems (McLean
et al., 1999; Sellers et al., 2002; Pack-Melton et al., 2006; Kiedaisch et al., 2006).
The general idea is to replace complex, maintenance-intensive, drag-producing
high-lift components, totally or partially, with simpler flaps and simple leading-
edge droops and then control separation by means of autonomous actuators.
A central challenge is the optimization of control to yield the maximum beneficial
results with the lowest associated cost. One approach to the problem is to introduce
perturbations at multiple locations or at multiple frequencies and then optimize the
control (e.g., Pack-Melton et al., 2004; Kiedaisch et al., 2006; Greenblatt, 2007).
A disadvantage of simplified high-lift systems is that deflection of the simple flaps
and slats reduces the effective chord length and consequently projected wing area,
particularly at large deflection angles d > 40 deg. In contrast, Fowler flaps and
leading-edge devices have precisely the opposite effect, and this intensifies the
challenge of achieving meaningful high lift by simple devices. The problem of
so-called simplified high lift, three-dimensional effects including sweep, as well as
the effects of Reynolds and Mach number, are discussed in detail in Chapter 8.

B. Effects of Streamwise Curvature


Streamwise curvature can have a profound effect on the efficacy of control, as
well as the range of optimum frequencies and amplitudes. Perturbations intro-
duced on a curved surface may be amplified by the Kelvin–Helmholtz and Görtler
mechanisms simultaneously, resulting in a corrugated spanwise vortex structure.
On airfoils, the leading-edge radius can place limitations on control strategies and
significantly affect the leading parameters. The relatively “simple,” classical problem
of a cylinder placed in crossflow presents significant challenges from an AFC
perspective.

1. The Coanda Cylinder


An example where streamwise vortices play a major role in the enhancement of
mixing between a wall jet and its surrounding fluid is the wall jet over a curved
surface. Wall jets in the absence of streamwise curvature bear some similarity to
the plane mixing layer, although they are much more complex. Their large coherent
DEVELOPMENT AND APPLICATION OF AFC 39

structures are generated by two dominant instability modes: an inflectional instability


in the outer region that resembles the mixing layer and a viscous instability near
the surface. A wall jet flowing over a circular cylinder has a unique ability to wrap
itself around the cylinder’s surface. The solid lines shown in Fig. 8a represent the
streamlines calculated from mean-velocity measurements. The radial distance
from the wall to the location at which the mean streamwise velocity component
has its maximum, Umax, is denoted by ym, and the distance from the wall at which
the velocity in the outer part of the flow is equal to _12 Umax is denoted by y2 and
shown in Fig. 8a as a dashed curve. It is customary to equate y2 with the width of
the jet. Both ym and y2 are plotted in Fig. 8b as functions of the distance from the
jet’s origin. The data are normalized by the slot width b.
As already mentioned, the curvature of the surface adds to the complexity of
this flow because it creates yet another instability mechanism, associated with
centrifugal forces, which is commonly referred to as the Görtler instability. In
order to assess the effects of the curvature, a flat plate was placed next to the
nozzle and results were compared under otherwise identical flow conditions. The
locations of y2 and ym of the curved jet are farther away from the wall than those
of the plane wall jet, indicating that the curved flow diverges more rapidly (Fig. 8b).
The normalized mean-velocity profiles in both flows were almost identical,
suggesting that the presence of the streamwise vortices does not distort the flow.
It was shown by Neuendorf and Wygnanski (1999) that the mean turbulent
intensity in the curved wall jet was much larger than in a corresponding plane wall
jet, even when both exit velocities and nozzle widths were identical. This observa-
tion was confirmed by Han et al. (2004) and is plotted in Fig. 9. The distributions
of the turbulent intensities emphasize the lack of equilibrium between the mean
velocity and the turbulence intensity in the curved wall jet compared with its plane
counterpart. Furthermore, the maximum relative intensities in the curved flow are
approximately 50% larger than in the plane flow and the locations of these maxima
are farther away from the solid surface than in the corresponding plane wall jet.
These differences were attributed to the presence of longitudinal vortices gener-
ated by the curvature. The existence of streamwise vortices was first observed by
flow visualization and was also inferred from spatial correlations of the spanwise
(z-direction) velocity component w, which served as a good indicator for the average
wavelength of these vortices. Furthermore, as the width of the flow increased with
increasing distance from the nozzle, the preferred wavelength increased as well,
scaling linearly with the thickness of the wall jet. Counter-rotating, streamwise,
vortex pairs were revealed by Neuendorf et al. (2004) using PIV in conjunction
with pattern-recognition techniques. This allowed the authors to overcome the
low-frequency random meander of the vortices and describe them quantitatively
in a statistical sense. The vortex meander provided an explanation for the almost
perfect two-dimensionality of the mean flow, in spite of the considerable strength
of the counter-rotating streamwise vortices.

2. Effect of Airfoil Leading-Edge Curvature


The effect of airfoil leading-edge curvature on separation control of an incom-
pressible flow can best be seen by comparing NACA 0012 (characterized by leading-
edge stall) and NACA 0015 (characterized by trailing-edge stall) airfoils, both
40 D. GREENBLATT AND I. J. WYGNANSKI

0 deg
a)

g
de
45
90 deg

13
5
de
g
180 deg

b) 50 LIVE GRAPH
Click here to view

ym /b, curved jet


40
y2/b, curved jet
ym /b, plane jet
y2/b, plane jet
ym /b, y2 /b

30

20

10

0
20 40 60 80 100 120 140 160
x /b

Fig. 8 Flow of a wall jet over a circular cylinders: a) streamlines and b) comparison of
relative widths of the flow with a straight wall jet (Neuendorf and Wygnanski, 1999).

equipped with zero-mass-flux blowing slots. The relatively large leading-edge radius
on the NACA 0015 (2.48% of chord) results in relatively gentle trailing-edge stall;
the smaller leading-edge radius of the NACA 0012 (1.58% of chord) leads to a
bubble-bursting mechanism that gives rise to alternating intervals of partial attach-
ment and separation, but with no regular frequency (McCroskey et al., 1982;
DEVELOPMENT AND APPLICATION OF AFC 41
LIVE GRAPH LIVE GRAPH
a) Click here to view x/b b) Click here to view
3.0 3.0 x/b
30.3 20
53.0 30
68.2 40
2.5 83.4 2.5 50
98.5 60
106.1 75
2.0 113.7 2.0 90
121.2 105
128.8 120
136.4
y/y2

1.5 1.5 135


144.0 150
151.6
159.1
1.0 166.7 1.0
174.3

0.5 0.5

0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.00 0.05 0.10 0.15 0.20
u'rms/Umax u'rms/Umax

Fig. 9 Turbulent intensity measured along the path of the flow: a) a curved wall jet
and b) a straight wall jet (Han et al., 2004).

Greenblatt and Wygnanski, 2003). Control on the NACA 0012, downstream of the
bubble, effectively limited its size to the location of the actuator, trapping it upstream.
The combination of high leading-edge curvature and excitation downstream of the
bubble rendered the flow less receptive to the excitation than on the NACA 0015
under otherwise similar conditions.
Basic differences in response to control are seen by comparing the dependence
of lift and moment coefficients on the reduced frequency F+ at relatively low-
amplitude excitation (·Cµ Ò  0.2%; Fig. 10). For the NACA 0012, modest lift
increments are produced over the range 0.5 < F+ < 5, with no clear optimum F+,
but increasing the frequency improved the pressure recovery (not shown), with
consequent reduction in nose-down moment coefficient. Higher amplitudes
(approximately ·Cµ Ò > 1%) at F+ < 2 were beneficial (not shown); this is believed
to be due to the large harmonic content associated with the decay of higher-
amplitude perturbations. In contrast, excitation on the NACA 0015 at 0.5 < F+ < 1
was effective in increasing lift, but totally ineffective at F+ > 2 owing to rapid
decay of the perturbations toward the trailing edge. For a given excitation fre-
quency, the NACA 0012 required ·Cµ Ò of the order of 0.1% to bring about
increases in post-stall lift (not Cl,max) and significant reductions in post-stall
unsteadiness. In contrast NACA 0015 had negligible post-stall unsteadiness asso-
ciated with it and required an order of magnitude less momentum input to attain
meaningful post-stall lift enhancements.

3. Circular Cylinder: The Bluff-Body Control


The crossflow over a nominally two-dimensional circular cylinder is a proto-
typical bluff-body problem, with direct and indirect applications to fixed-
and rotary-wing aircraft, wind-exposed buildings, and underwater structures.
42 D. GREENBLATT AND I. J. WYGNANSKI

a) 0.35
Re=200,000, Cm=0.2%
LIVE GRAPH
Click here to view NACA 0012: a=as+2°
0.25

DCl NACA 0015: a=as+4°

0.15

0.05

–0.05

b)
0

–0.02

Cm
–0.04
NACA 0012: a=as+2°
NACA 0015: a=as+4°
–0.06
Re=200,000, Cm=0.2%

–0.08
0 1 2 3 4 5 6
F+

Fig. 10 Effect of leading-edge radius on the optimum reduced frequency required


for a) lift and b) moment control (Greenblatt and Wygnanski, 2003).

Controlling this flow poses substantial challenges because the separation line is not
precisely (geometrically) determined (subcritical: 70 deg < bs < 85 deg; supercritical:
120 deg < bs < 140 deg; see Fig. 11a) and multiple instabilities are present simulta-
neously, namely viscous (Tollmien–Schlichting; see Schlichting, 1979) instability
of the laminar boundary layers, centrifugal (Görtler) instabilities due to curvature
(see Saric, 1994), inflectional (Kelvin–Helmholtz, see Michalke, 1964, 1965)
instability of the separated shear layer, and the ubiquitous global instability that
drives vortex shedding (Huerre and Monkewitz, 1990).
Active control from a slot has been attempted by many investigators (e.g.,
Hsiao et al., 1990; Williams et al., 1991; Schewe, 1983; Pal and Sinha, 1997;
Heine et al., 1997; Amitay et al., 1998; Liu and Brodie, 1999; Béra et al., 2000),
producing profound effects on transition and separation, with consequences for
lift, drag and vortex shedding. For example, Hsiao et al. (1990) showed that a
DEVELOPMENT AND APPLICATION OF AFC 43

suction peak could be generated on one side, resulting in Cl = 0.6 with the exci-
tation location at 100 deg when F+  fe D/U•  1; similar observations were
made for F+ = 1.5 by Amitay et al. (1998), who also showed that the separation
point could be moved by approximately 60 deg, resulting in a 25% reduction in
form drag.
At subcritical Reynolds numbers, Naim et al. (2007) observed that separation is
controlled by two distinctly different mechanisms: by forcing laminar–turbulent
transition when applied at relatively small angles (30–60 deg) from the forward
stagnation point; and by directly forcing the separated shear layer at larger angles
(see the configuration in Fig. 11a). They further observed that the universal
Strouhal law also holds for active control on cylinders, as long as the excitation
frequency is significantly higher than the natural vortex-shedding frequency

LIVE GRAPH
Click here to view

Fig. 11 Nondimensional shedding frequency based on two lengthscales that charac-


terize the wake thickness (90 deg £ b £ 105 deg). a) Symbol convention with respect to
the cylinder; b) nondimensional shedding (Naim et al., 2007).
44 D. GREENBLATT AND I. J. WYGNANSKI

(Fig. 11b). This extension of the universal Strouhal law is attributed to separation
control, which narrows the wake, thereby increasing the vortex shedding frequency
(VSF) and decreasing the drag. However, drag-reduction predictions based on the
related model (Roshko, 1954a, b) are too small (not shown), mainly because the
model does not account for partial attachment of the separated shear layer that
results from active flow control.
At specific forcing angles, form drag can be either increased or decreased by
altering the modulating frequency (Fig. 12). Lock-in, by means of the low-
frequency burst-modulated excitation at F+ = 0.16, increased the pressure drag by
32% (b = 55 deg) and 17% (b = -55 deg). Roshko’s (1954a, b) model (see Naim
et al., 2007) predicts a 16% pressure drag rise due to lock-in at this frequency. At
the higher frequency (F+ = 0.52), at the same Cµ and b, the form drag decreases by
16% and decreases even further at higher b. This higher frequency is outside the
range of possible “lock-in” frequencies, although it is possible to increase the
upper bound of the “lock-in” region by significantly increasing the amplitude of
the excitation (e.g., Griffin, 1989).

C. Electro- and Magnetohydrodynamic Actuation


Recent years have witnessed a trend toward the use of electro- and magneto-
hydrodynamic active separation control. An example of the latter (wall-parallel
Lorentz force actuation) is achieved in weakly conductive fluids, such as sea
water, by arranging alternating strips of flush-mounted electrodes and permanent

LIVE GRAPH
Click here to view

Fig. 12 Form-drag manipulation using frequency locking at frequencies close to the


shedding frequency and separation control frequencies (burst mode excitation,
ReD = 100,000) (Naim et al., 2007).
DEVELOPMENT AND APPLICATION OF AFC 45

magnets. Although these actuators have long been investigated for drag reduction
of turbulent boundary layers, they have only recently been exploited within the
context of active separation control. The unique aspect of these actuators is that
virtually any Lorentz force distribution can be produced by suitably arranging the
magnets and electrodes (Weier and Gerbeth, 2004). Performance increments are
of a similar order to those achieved by means of zero-mass-flux blowing.
Significant research effort has recently been expended on plasma actuators
(e.g., Post and Corke, 2004, 2006), with particular emphasis on so-called dielec-
tric barrier discharge (DBD) actuators, which involve the asymmetric arrange-
ment of electrodes separated by a dielectric layer. The actuator is attractive for
separation control applications because it is surface-mounted (not requiring
internal volume), it is a simple design with no moving parts, it operates nominally
along a line, and it is electrically efficient. Commonly used actuators (of order
10 Vptp and 10 W), however, appear to have somewhat limited control author-
ity (see the discussion in the next paragraph). Research has advanced in two
main directions: 1) understanding the physics of the momentum coupling into
neutral air (e.g., Enloe et al., 2006) and 2) application to active flow control, in
particular active separation control (e.g., Post and Corke, 2004, 2006); here we
focus on the latter.
Initial investigations with application to active separation control were per-
formed by Post and Corke (2004, 2006), who employed a DBD actuator in
“steady” and pulsed modes at conventional low speed and low Reynolds numbers
(Re > 200,000). These investigations revealed a number of important characteris-
tics and limitations regarding the technique: plasma actuation in steady and pulsed
modes has an effect in the post-stall regime, but produces a small effect on key
aerodynamic coefficients (e.g., DCl,max  0.2 and changes to (l/d)max are negli-
gible); plasma actuation effects deteriorate with increasing Re; pulsed-mode
actuation is superior to steady actuation and, for increasing Re, steady plasma
control can have a slightly deleterious effect on DCl,max. Similar effects were
observed previously, for example, using piezoelectric zero-mass-flux jet actuators
(Margalit et al., 2002). With pulsed mode actuation, the minimum voltage (pre-
sumed to be proportional to momentum) required to attach an otherwise separated
flow occurred at F+ slightly larger than 1, consistent with the data of Nishri and
Wygnanski (1998). At lower Reynolds numbers, the proportion of actuator
momentum, and therefore control authority, increases. This observation was
exploited by Göksel et al. (2006, 2007) to perform a series of experiments in the
range 3000 £ Re £ 80,000, discussed later in this chapter.

D. Unsteady Separation and Attachment


Up to now, we have dealt with time-invariant active control; that is, the flow
remains attached in a time-mean sense. The process by which the flow separates
from, or attaches to, a surface is important when we want to control a process
whose characteristic timescale is much larger than the typical period of eddy
passage when periodic actuation is applied: O(Xsep /U•). Typical examples include
the response of vehicle control-surface flow (Amitay et al., 2004), intermittent
dynamic stall control (Greenblatt et al., 2001), and control of wake vortices
(Greenblatt et al., 2006a, b).
46 D. GREENBLATT AND I. J. WYGNANSKI

1. Essentials of Forced Separation and Attachment


The forced reattachment of an initially separated flow was studied by abruptly
changing the frequency and amplitude of periodic perturbations emanating from
a slot at the flap shoulder (Darabi and Wygnanski, 2004a). The minimum reattach-
ment time occurred at an optimal excitation frequency of F+  1.5 (Fig. 13a),
which was independent of amplitude and flap inclination. The timescales of the
excitation are at least an order of magnitude smaller than the typical reattachment
LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view

Fig. 13 a) Reattachment time tr as a function of reduced frequency F+ for different


forcing amplitudes: Ddr = 6 deg; ReL = 124,000 (Darabi and Wygnanski, 2004a) and
b) Separation time in excess of the uncontrolled process, ts – ts0 as a function of a
combined forcing parameter (Darabi and Wygnanski, 2004b).
DEVELOPMENT AND APPLICATION OF AFC 47

times. The reverse process, forced separation, was also studied by abruptly chang-
ing the slot excitation (Darabi and Wygnanski, 2004b). A complete cessation of
the actuation resulted in the formation of a large vortex above the flap akin to the
familiar dynamic stall vortex (DSV) seen over oscillating airfoils in pitch. The
DSV temporarily increased the aerodynamic load over the flap, before it dropped
to its low separated value. The duration of this overload decreased as the flap
inclination increased. The use of periodic excitation during separation slowed
down the rate of separation and changed its character, depending on the amplitude
and frequency used. Forcing separation by switching the excitation to a high fre-
quency (3 £ F+ £ 8) reduced or even eliminated the increase in flap loading that is
associated with the DSV. A switch to low frequencies (F+ < 1) extended the dura-
tion of separation and increased the transient overload during the initial stage of
the process. The time to complete separation correlated well with the empirically
determined parameter ·CmÒ exp(-F+) (Fig. 13b). The switch to a low frequency or
the reduction in amplitude at low frequencies (F+ < 2) resulted in the creation of a
DSV, whereas the switch to a high frequency resulted in separation progressing
upstream from the trailing edge. It may be useful to extend this technique to airfoil
flows, where the dynamic stalling characteristics may be controlled, producing
either leading- or trailing-edge stall.

2. Incompressible Dynamic Stall Control


Dynamic stall on rotorcraft retreating blades results in dramatic loss of lift and
large pitching moments, which transmit excessive and damaging impulsive loads
to the flight control system and airframe (Lorber et al., 2000). A dominant feature
is the DSV, mentioned in the preceding section and often referred to as “bubble
bursting,” which is generated when the blade pitches at a sufficiently high pitch rate
beyond its static-stall angle (Carr, 1988). Although dynamic stall occurs in practice
near the tips of rotor blades, where compressibility effects cannot be ignored
(Ma  0.4), dynamic stall is seen to originate and persist in regions inboard, closer
to the hub, where the flow can be considered to be incompressible (Bousman,
1998). Thus, the study of dynamic stall control in both regimes is justified.
The definition of dynamic stall control is prone to some subjectivity, because it is
ultimately dictated by practical considerations such as allowable design loads. When
an airfoil pitches dynamically into the post-stall regime (a max > as), large lift and
large negative pitching moment or moment excursions (Cm,min or Cm,exc) result.
Bousman (2000) developed so-called dynamic stall functions Cl,max = f(Cm,min) and
Cl,max = f(Cdp,max) for the evaluation of dynamic stall control techniques. A different
metric was proposed that requires increasing or maintaining Cl,max while constraining
Cm to be commensurate with the pre-stall excursions, or so-called allowable excur-
sions, Cm, A (Greenblatt and Wygnanski, 2001a).
A comparison of two prototypical airfoils (leading- and trailing-edge stallers:
NACA 0012 and NACA 0015, respectively) is considered here, with control being
produced by means of zero mass-flux blowing (Greenblatt, 2006b). The dynamically
pitching NACA 0012 generates lift beyond as, after which dCl/da increases, and
this coincides with the onset of moment stall (see Carr, 1998), which is abrupt and
relatively severe (resulting in large Cm,exe) (Figs. 14a and 14c). In contrast, the
dynamically pitching NACA 0015 does not significantly increase lift beyond
48 D. GREENBLATT AND I. J. WYGNANSKI

LIVE GRAPH
LIVE GRAPH Click here to view
Click here to view

LIVE GRAPH
LIVE GRAPH Click here to view
Click here to view

Fig. 14 Comparison of NACA 0012 (Re = 240,000) and 0015 (Re = 300,000) dynamic
stall control at F+ = 0.6 (Greenblatt, 2006a).

as, and dCl/da decreases. This is accompanied by relatively gentle moment stall
with relatively small excursions (Figs. 14b and 14d). In the former case, moment
stall is significantly more severe, typically requiring larger momentum coefficients
to effect control, and this was achieved by “trapping” the bubble upstream of the
forcing slot. The minimum ·Cµ Ò required to effectively control dynamic stall was
observed to be proportional to (a max – as)2 (not shown). For the NACA 0015,
relatively low reduced frequencies (F+ < 1) were capable of exerting control
at relatively modest forcing amplitudes (0.1% £ ·Cµ Ò £ 0.5%). The example
illustrated in Figs. 14b and 14d, with forcing at F+ = 0.6 and ·Cµ Ò = 0.21, shows a
DEVELOPMENT AND APPLICATION OF AFC 49

simultaneous increase in Cl,max, with virtual elimination of hysteresis and a reduction


in Cm,exc < Cm,A.
A summary of Cl,max and Cm,exc data for different F+ values as functions of Cµ
showed that Cm,exc exhibited a strong sensitivity to F+ for both airfoils (not shown).
For leading-edge stall, Cµ produces a direct decrease in Cm,exc irrespective of F+,
such as when Cm exceeds some threshold, it is seen that Cm,exc µ ln(1/·Cµ Ò). In fact,
F+ = 0.6 and 1.1, which were observed to be the most effective values for NACA
0015 control (i.e., requiring the lowest ·Cµ Ò), are the least effective for NACA
0012 control: this is consistent with static data, (Greenblatt and Wygnanski, 2003)
This retrospectively explains the relatively large Cµ required for effective control
at F+ = 0.6 shown in Figs. 14a and 14c. The most effective reduced frequency is
F+ = 3.5 and requires approximately 4 times less ·Cµ Ò than at F+ = 0.6 to render
Cm,exc < Cm,A and up to 30 times less Cµ in order bring about a meaningful reduction
in Cm,exc. Unfortunately, the degree to which forcing reduces Cm,exc on the NACA
0015 varies in a non proportional manner that is dependent on F+.

3. Control at Flight Conditions


Scaling AFC up to rotorcraft flight conditions presents some significant
challenges, because amplitude and frequency must be scaled up proportionately,
and this poses a significant challenge for actuator development. In addition, com-
pressibility is known to have a deleterious effect on steady AFC at around Ma > 0.7
(static airfoil control is discussed in Sec. IV.2). Carr (1988) pointed out that, for
dynamically pitching airfoils, compressibility can have a profound effect on
dynamic stall, even at relatively moderate Mach numbers (i.e., Ma = 0.3), when
the flow can be supersonic in the leading-edge region, and this intensifies the
challenge to effective AFC (Carr and Chandrasekhara, 1996). When attempting
dynamic stall control, compressibility clearly becomes a factor, because typical
full-scale Mach numbers on a rotorcraft retreating rotor blade in the vicinity of
dynamic stall are in the range 0.3–0.5. The effect of Reynolds number is somewhat
less understood, owing to the difficulty of varying Reynolds number significantly
without introducing compressibility effects.
Nagib et al. (2001, 2007) studied control using zero-mass-flux jets introduced
from spanwise slots at various locations on the upper surfaces of oscillating VR-7
and VR-22 airfoil models at 0.1 £ Ma £ 0.4. Control was demonstrated over a
wide range of mean angles of attack of the oscillating airfoil, from light to deep
stall conditions. Comparable modifications of the aerodynamic coefficients were
observed throughout this Mach-number range, proving that F+ and ·Cµ Ò were
maintained, even in the presence of local shocks. Therefore, it appears that zero-
mass-flux blowing is limited only by the ability to generate the adequate forcing
conditions at the higher Mach numbers required for rotorcraft applications.

E. Very Low Flight Reynolds Numbers


The traditional trend in aerodynamics has been toward faster and larger vehicles
(i.e., an increase in Re), but recent years have witnessed the opposite trend, with a
demand for unmanned vehicles of successively decreasing dimensions and flight
speeds. Two main classes of unmanned vehicle have emerged: so-called micro aerial
50 D. GREENBLATT AND I. J. WYGNANSKI

vehicles (MAVs: 7.5 cm £ b £ 15 cm (e.g., Mueller, 1999) and nano air vehicles
(NAVs: typically b < 7.5 cm), although these definitions vary somewhat. Achieving
sustained flight of these vehicles brings significant challenges, because the combi-
nation of small dimensions and low flight speeds, particularly during loiter, results
in Re < 70,000, where the assumptions associated with conventional low Reynolds
number aerodynamics begin to lose validity (e.g., Carmichael, 1981).
Dramatic improvements in performance were observed on a 10.9 cm chord,
17% thick airfoil at Reynolds numbers of 50,000 and 30,000, where zero-mass-flux
perturbations were produced from a slot at 2% chord (Greenblatt and Wygnanski,
2001). On small-scale vehicles, internal volume will limit the size of actuators.
An alternative is to used DBD actuators, as these are surface-mounted and do not
require internal volume (e.g., Corke et al., 2004). Although these actuators are
typically driven in the kilohertz frequency range, they can be pulsed at lower
(separation control) frequencies, resulting in significant improvements to Cl,max,
which increase with reductions in Re (Greenblatt et al., 2008; Mabe et al., 2009).
For low MAV Reynolds numbers (Re  20,000), modulation increased Cl,max by
more than a factor of 2 on an Eppler E338 airfoil, and typical low-Re hysteresis
was eliminated. Of particular interest from an applications perspective was that
performance, measured here by Cl,max, did not decrease with decreasing duty cycle
(DC), and hence power input. In fact, duty cycles of around 0.66% were sufficient
for effective separation control, corresponding to power inputs of about 1.2 mW
per centimeter of the wingspan (Göksel et al., 2006). The details of profile design
were seen to have a significant effect on airfoil performance and the optimum F+
range (Figs. 15 and 16).

VII. Wake Vortex Control


There are many technological challenges posed by trailing vortex wakes, and
we consider only elements of controlling vortices trailing large aircraft (e.g.,
Rossow, 1999; Spalart, 1998) and those found on rotary-wing aircraft (e.g.,
Leishman, 2000). The active concept most commonly advocated to reduce the
wake hazard is to force instabilities in trailing vortices by oscillating control sur-
faces periodically such that the integrated aerodynamic loads do not fluctuate.
This accelerates unstable vortex growth such that the vortices ultimately interact,
pinch-off, and degenerate into harmless small-scale turbulence (Chevalier, 1973;
Crouch, 1997; Crow, 1970; Crow and Bate, 1976). On rotorcraft, a major source
of noise and vibration arises from a rotor blade cutting through the tip vortex shed
by its predecessor (blade-vortex interaction, BVI). Active methods seek to increase
the “miss distance” between the rotor blade and tip vortex or diffuse the vortex,
for example using trailing-edge flaps. In contrast to active methods that rely on
deflection of conventional control surfaces, AFC seeks to manipulate the vortices
by controlling the vortex sheet locally. Here we examine AFC as a means of con-
trolling the primary characteristics of vortices, namely their location (centroids),
strength, size, and associated velocity components.

A. Basics of Vortex Perturbation


A simple, but powerful, method for predicting vortex characteristics is the
method of Betz (1932), as modified by Donaldson (1974), which does not explicitly
DEVELOPMENT AND APPLICATION OF AFC 51

a) LIVE GRAPH
Click here to view Re = 50,000, DC = 3%, α = 18°, [Cm] ≈ 0.001%
Re = 35,000, DC = 5%, α = 18°, [Cm] ≈ 0.01%
0.8 Re = 20,500, DC = 5%, α = 18°, [Cm] ≈ 0.2%
Re = 20,500, DC = 5%, α = 14°, [Cm] ≈ 0.2%

0.6

∆Cl

0.4

0.2

0
0 2 4 6 F+ 8 10 12

b) LIVE GRAPH
Click here to view
0.8
α = 20°, Re = 3000, [Cm] ≈ 2.0%

α = 15°, Re = 6,000, [Cm] ≈ 0.5%


0.6 α = 20°, Re = 6,000, [Cm] ≈ 0.5%

α = 20°, Re = 9,000, [Cm] ≈ 0.2%


∆Cl

0.4

0.2

0
0 1 2 3 4 5
F+

Fig. 15 Control using DBD plasma actuation on a) an Eppler E338 airfoil and b) a flat
plate, showing post-stall lift dependence on reduced frequency (Greenblatt et al., 2008).

treat the rollup mechanism, but rather employs three conservation relations between
the span loading G( y) and the rolling-up vortex G ¢(r). Despite the relative simplic-
ity of the method, and some well-known limitations (Widnall, 1975), predictions
are in good agreement with aircraft-wake wing tip and flap vortex measurements
(Donaldson et al., 1974). The method therefore allows rapid assessment of the
viability of a particular AFC technique. On approach for landing, where flap vorti-
ces dominate the wake, manipulating the bound circulation G locally modifies the
52 D. GREENBLATT AND I. J. WYGNANSKI

Fig. 16 Smoke-flow visualization illustrating the effect of DBD pulsed plasma con-
trol on an Eppler E338 airfoil (Re = 20,000) and a flat-plate airfoil (Re = 3000),
(Greenblatt et al., 2007): a) E338 baseline, b) E338 control at F+ = 1.0, duty cycle = 5%,
c) flat plate baseline, and d) flat plate at F+ = 0.42, duty cycle = 5%.

vortex sheet strength (or shed vorticity) g = dG/dy. The local control of separation, or
circulation, impacts directly on the location and formation of individual vortices.
The basic concept of localized control of the vortex sheet, and hence control over
individual vortices, was investigated by Greenblatt et al. (2006). The upper parts of
Fig. 17 show schematics of the experimental setup, consisting of a semispan wing
model with an inboard flap consisting of control slots. Here, both passive and active
control (Figs. 17a and 17b) were applied over finite segments of the flap, namely
inboard and outboard, and the resulting span-load distribution and trailing-edge
pressures are shown in the figures. In both instances, the measured wing lift and
pitching moments were the same, and, in order to dynamically perturb the vortices,
control would be applied alternately, in a time-dependent manner. Seven-hole probe
measurements were made in the wake of the wing flap corresponding to the four
conditions. For purposes of comparison with theory, the measured span-load distri-
bution was used to compute the vortex characteristics based on the method of
Donaldson (1974) and Betz (1932) already mentioned (see Table 1).
Table 1 shows that main vortex characteristics trends are qualitatively predicted,
and, for all cases examined by Greenblatt et al. (2002a), the centroid movement, as a
fraction of semispan s, was well predicted. An important observation, consistent with
theory, is that control over the vortex sheet in the vicinity of the flap edge, particularly
the location of (dG/dy)max, produced a direct effect on the vortex centroid. Thus,
when vortex control is decoupled from the conventional control surfaces, changes to
wing lift do not necessarily correlate with vortex-centroid control authority.

B. Axial Flow Control Using the “Bernoulli Effect”


Axial velocity in vortices is either wake-like (toward the vehicle) or jet-like
(away from the vehicle). Axial velocity has profound effects on vortex stability,
and periodic changing from wake-like to jet-like flow may precipitate vortex
bursting (Chevalier, 1973). This jet/wake phenomenon can be termed the
LIVE GRAPH DEVELOPMENT AND APPLICATION OF AFC 53
Click here to view
Passive Passive Semispan wing Active Active Semispan wing
inboard control outboard control inboard control outboard control

Flap Flap
a) b)
1.6 –0.4 1.6 –0.4
Cp,te C p,te

1.2 0.2 1.2 0.2

Cl Cl

0.8 0.8
Inboard passive control Inboard active control
Outboard passive control Outboard active control
0.4 0.4
Re ≈ 500,000, α = 8 deg Re ≈ 500,000, α = 8 deg
(δi,δo,δt) = (20 deg, 20 deg, 0 deg) (δi,δo,δt) = (20 deg, 20 deg, 0 deg)

0 0
0 0.33 0.67 1.00 0 0.33 0.67 1.00
y/s y/s

Fig. 17 Control of span loading to manage trailing vortices (Greenblatt et al., 2006):
a) passive and b) active control.
LIVE GRAPH
Click here to view
“Bernoulli-effect” (Batchelor, 1964; Spalart, 1998), because application of the
equation to the streamline upstream of the wing and through the vortex center
explains the phenomenon. In this context, Bernoulli’s equation for the axial
velocity along the vortex center can be written in cylindrical coordinates as
2
ÈVx (0) ˘ k (G¢ / r1 ) 2 r g DH
Í ˙ = 1 + -
Î U• ˚ q• q• (3)
  
pressure drop viscous head drop
in the vortex in the vortex

where the pressure drop along the vortex is proportional to (G ¢/r1)2 (Green, 1995)
and k is an positive constant. The pressure-drop term always acts to produce a

Table 1 Comparison of vortex changes based on inviscid rollup relation


predictions and near-wake measurements

Partial flap: passive control Partial flap: active control

Experiment Theory Experiment Theory

Centroid movement, %s 1.30 1.50 0.82 0.78


Vortex strength ratio 1.23 1.05 1.17 1.27
Peak velocity ratio 1.30 1.29 1.38 1.45
Inner size ratio, r1 0.71 — 0.62 —
Outer size ratio, r2 — 0.82 — 0.88
54 D. GREENBLATT AND I. J. WYGNANSKI

jet-type flow and the head-drop term always acts to produce a wake-type flow.
Thus the axial flow direction will be determined by the relative magnitudes of
these two terms and, in the absence of AFC, depends on the wing loading. The
azimuthal velocity, or vortex diffusiveness, is also affected and is proportional to
[Vx (0)/U•]1/2.
Greenblatt et al. (2005) exploited both terms on the right-hand side of Eq. (3)
by controlling the separation flow over a deflected tip flap. Axial flow over the
flap, axial flow in the vortex, and axial vorticity with corresponding centroid loca-
tions are shown in Figs. 18a–c, 18d–f, and 18g–i, respectively (two-dimensional
and stereo PIV data). In the baseline state, the relatively large wake associated
with the separated shear layer (Fig. 18a), is rolled up into the vortex, producing
relatively large DH and r1 (Fig. 18g) and hence a minimum velocity deficit
U/U•  0.6 (Fig. 18d). Low-amplitude control reduced the shear-layer losses
(Fig. 18b) and thereby virtually eliminated the velocity deficit (Fig. 18e) by reducing
both DH and r1 (Fig. 18h). These quantities are further reduced with higher-
amplitude control (Fig. 18c), bringing about a dramatic increase in Vx (0)/U•
(Fig. 18f). Unfortunately, the higher rotational velocities Vq near the vortex center
LIVE GRAPH
Click here to view

Fig. 18 Control of a wing-tip vortex via tip flap separation control: a–c) axial flow
over the tip flap, d–f) axial velocity in the vortex at x/c = 2, and g–i) axial vorticity in
the vortex at x/c = 2 (Greenblatt et al., 2005).
DEVELOPMENT AND APPLICATION OF AFC 55

centrifuge the PIV seed particles out of the vortex core, but an independent seven-
hole probe showed that Vx(0)/U• > 1.6. With less of the separated flow rolling up
into the vortex, the vortex becomes less diffuse and its centroid moves laterally
1.5% and longitudinally 3.5% of the chord length.

C. Dynamic Vortex Perturbations


Williamson et al. (1998) identified three distinct instability length scales trailing
a wing in a towing tank studies: a small-scale “braid wake” that scales with the thick-
ness of the two shear layers separating from the upper and lower surfaces of the
wing trailing edge; a short-wave instability that scales with the primary vortex core
dimensions; and a long-wave instability far downstream representing the classical
“Crow” instability (Crow, 1970), scaling with the distance between the two primary
vortices. The coexistence of short- and long-wavelength instabilities was also
observed at low Reynolds numbers, and similar observations have been made in the
wakes of commercial airlines (Fig. 19). Separation control is considered here as a
means to perturb or promote instabilities at these vastly different wavelengths.

Fig. 19 Successive photographs of an airline trailing-vortex pair, taken from the


ground with a telephoto lens. Reproduced courtesy of O. Savas (Bristol et al., 2004).
56 D. GREENBLATT AND I. J. WYGNANSKI
LIVE GRAPH
Click here to view
103
Decreasing AR,
increasing Lf /c
102

10
Increasing AR
decreasing Lf /c
λ/b 1
AR = 4; Lf /c = 0.3

10–1

Dynamic seperation
10–2 control regime
Quasi-steady or
separation control shear-layer Effective
10–3 control separation Separation control
control less effective

10–4
10–4 10–3 10–2 10–1 1 10 102
+
F
Fig. 20 Relationship between vortex wavelength and AFC reduced frequency.

To estimate instability wavelengths in the context of separation control, we


express the wavelength as a fraction of wingspan, l /b = U• / fb, and define the
ratio e = Xte /c. Using the definition of F+, we can then write

l/b = e /F+ AR (4)

and plot l /b versus F+ for various values of e and AR (Fig. 20). The solid line
represents a “typical” configuration (e = 0.25 and AR = 8) and the top hatched line
represents the configuration tested by Greenblatt et al. (2005) (e = 0.3 and AR = 4).
Note that the line moves up with increasing e and decreasing AR, and vice versa.
Note from Fig. 20 that the high end of the effective separation; control frequencies
(approximately F+  3) correspond to wavelengths g /b << 1. This is typical of the
“braid” instabilities identified by Williamson et al. (1998). The opposite end of the
effective separation-control spectrum, F+  0.3, corresponds approximately to
the lower end of the short-wave instability. For F+ < 0.3, separation control begins
to lose its effectiveness and the boundary layer begins to stall dynamically.
Nevertheless, if higher-frequency (F+  1) separation control perturbations are
modulated at a lower frequency, then the shear layer can be forced to dynamically
attach and separate at F+ < 0.1. This would still correspond to the short-wave
instability identified by Williamson with g /b = O(1), but can also correspond to
longer-wavelength instabilities such as those of Crouch and Crow with g /b > 1.
At even lower F+, high-frequency separation control can be modulated in a quasi-
steady manner to achieve arbitrary long-wave vortex perturbations, corresponding
DEVELOPMENT AND APPLICATION OF AFC 57

to g /b >> 1. At the other end of the frequency spectrum, when perturbations are
introduced typically at F+ > 3, separation control begins to lose effectiveness,
boundary-layer control is not achieved, and the perturbations dissipate down-
stream. This corresponds to l /c <<1 and is not of obvious importance from a vortex
perturbation standpoint.

VIII. Summary
Modern AFC has emerged from laboratory-scale demonstrations in shear
layers, jets, wakes, and airfoils to full-scale ground-based demonstrations and
flight tests. Recognition that the control of large coherent structures can be
exploited to obtain a viable industrial benefit has been at the heart of these devel-
opments. As such, great potential exists for the effective control of mixing, jet
flows, and separation across wide Reynolds-number and Mach-number ranges.
Triple decomposition of controlled turbulent flows, combined with hydrodynamic
stability concepts (see Chapter 4), may provide the first step in developing rational
active flow control models.
Actuation methods that hold great promise are the exploitation of multiple-
mode and multiple-location control and three-dimensional perturbation as a means
of exciting streamwise instabilities. Research areas with potentially large payoffs
include simplified high-lift systems, three-dimensional configurations, and vortex-
dominated flows. The development of effective actuation remains a central
challenge (see Chapter 6), particularly for large-amplitude perturbations, and is
a critical element in the ultimate transitioning of AFC to industrial applications.
Finally, the optimization of surface shape to yield most effective interaction with
a prescribed method of active flow control is yet to be determined.
In this chapter, we have summarized our understanding of the physical concepts
underlying flows of engineering interest from an AFC perspective. The next chap-
ter will deal with AFC terminology followed by an in-depth discussion of flow
instabilities and closed loop control.
Chapter 3

Flow Control Terminology

Douglas G. MacMynowski*
California Institute of Technology, Pasadena, California
and
David R. Williams†
Illinois Institute of Technology, Chicago, Illinois

I. Introduction
Active flow control (AFC) is a multidisciplinary field involving the integration
of fluid mechanics with control in particular, and also with actuation and sensing
technology. This integration of different communities leads to the potential for
communication difficulties. In particular, many fluid dynamics researchers may
not be familiar with the terminology used by researchers in control. In this chapter,
we provide definitions (and recommend definitions where ambiguous) for much
of the control-related terminology not only in this book, but also currently being
used in the flow control literature. We hope that this will help reduce confusion by
providing some guidance on recommended terminology, resulting in more consis-
tency in usage, and helping to bridge some of the communication gap between the
fluid dynamics and control communities.
We begin by very briefly reviewing some standard control terminology; this can
be found in any control textbook, but is discussed here in the context of flow
control. We then discuss the terminology of flow control itself, and in particular
give a useful hierarchy for categorizing flow control ideas. Finally, some brief
comments on what is not flow control are given.

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Senior Research Associate, Control and Dynamical Systems. Senior Member AIAA.
†Professor, Mechanical and Aerospace Engineering. Associate Fellow AIAA.

The terminology discussion originated during AIAA Architectures & Algorithms in Flow Control
working group meetings from June 2002 to Jan. 2005 with input from A. Banaszuk, T. Bewley, L. Cattafesta,
S. Collis, T. Colonius, C. Rowley, and A. Seifert. The original document was endorsed by participants
at the First International Closed-Loop Flow Control Workshop, Jackson Hole, Wyoming, July 2005.

59
60 D. G. MACMYNOWSKI AND D. R. WILLIAMS

Some definitions associated with flow control are unambiguous (although their
usage in the literature may not be consistent). Others (such as the distinction
between “active” and “passive” control) do not have a precise definition or a
consistent usage in the literature. In these cases, we have attempted to articulate
the rationale that can be applied and we describe the most common usage; this
section should be used as guidance to improve understanding, not as a formal
definition of these terms.

II. Control Definitions


We begin by introducing many definitions commonly used in the control com-
munity that may not be familiar to researchers outside of this community. A more
detailed description of many of the concepts from dynamics and control that are
relevant to flow control can be found in Chapter 5. A useful and accessible text-
book on control is Åström and Murray (2008).
The first question to address, which is relevant to many of the flow-control-
specific definitions, is what defines the system being controlled. In the control
community, this is frequently referred to as “the plant”; this is just the system
under control. In flow control, the system is of course the fluid itself, but it may
also include, for example, actuator and sensor dynamics. There is, of course,
considerable judgment and insight involved in choosing the conceptual boundar-
ies of the system under control (both spatial and temporal), and what elements of
the physics are considered part of the system. Different flow control approaches
may appear to be categorized differently depending on how one draws these
boundaries, as discussed in Sec. III.F. Thus, for example, is the system under
consideration an aircraft, the inlet on that aircraft, or a separation somewhere
inside the inlet?
One of the most important distinctions in control is that between feedback or
closed-loop control (roughly synonymous) and open-loop control. We will defer a
more detailed discussion of this distinction, including diagrams comparing the
two types of control, until Secs. III.C. and III.D, after introducing some other
control terminology. Here we briefly introduce the concept, because many flow
control systems are open-loop, whereas almost all control theory is about closed-
loop control, and therefore much control terminology may be unfamiliar even to
those with significant experience in “flow control.”
There is feedback when two systems are connected so that the output of one
system affects the other. Thus, feedback control implies that the control signal
depends on information from the system that in turn depends on the control (this
is the “loop” referred to in the term “closed-loop”). Open-loop control may use
information from the system, but that information is not influenced by the opera-
tion of the control, so there is thus no information to determine how well the
control is working. Feedback is useful in dealing with uncertainty in knowledge
of either the system or disturbances (see Chapter 5). Furthermore, for linear sys-
tems, the dynamics of the system cannot be changed with open-loop control, that
is, without a feedback measurement from the plant. The output may change, for
example by superposition of a canceling wave, but the stability and dynamics do
not change. For this reason, many control researchers may not even be used to
considering open-loop forcing to be control. However, many flow control problems
FLOW CONTROL TERMINOLOGY 61

are interesting precisely because this is not true for nonlinear systems. The canoni-
cal example of open-loop forcing modifying the dynamics of a nonlinear system
is the ability to stabilize an inverted pendulum through periodic vertical excitation
of the base at an appropriate frequency. Flow control examples where periodic
excitation modifies the dynamics include separation dynamics and cavity flow
dynamics. A significant body of flow control work involves open-loop control, in
contrast to many other applications of control.
This also illustrates that nonlinearities may play a critical role in many flow
control problems. Nonetheless, much of control theory has been developed for
linear systems. Key concepts are briefly summarized below, starting with model-
ing and input/output notation, and then turning to control design.
Any active control (see, in contrast, the definition of passive flow control later
in this chapter) must have some means of influencing the system being controlled,
and the device(s) or component(s) that do so are the actuators. The actuator is
generally taken to be the entire subsystem that is added for the purpose of influ-
encing the system; for example, in the case of a synthetic jet being used in flow
control, one would typically consider the entire synthetic jet to be the actuator, not
just the voice-coil or piezoelectric transducer that powers the device.
The signals to the actuators are the control inputs (these are the inputs to the
system, which are the outputs of the controller itself). The vector of control inputs
is almost universally given the symbol u in the control community. In fluids work,
u is generally a velocity, and hence in this book we will typically use f for the
control input. However, researchers should expect to see different notation in the
controls literature. Note that the relevant control input variable(s) are the variables
being modulated, which may be the frequency or amplitude of a signal, for example,
rather than the actual voltage itself. In addition to the control input f, the system
is in general also influenced by disturbances, for which we use the symbol d.
The system may also have one or more sensors that measure some variables,
and these are typically given the symbol y in the control community. Although y
is also used frequently to denote a spatial coordinate, the distinction should be
clear from the context, and we will continue to use y here for sensor measure-
ments. In addition to the measured output variables, a model of the system being
controlled may also include additional output variables z whose magnitude is of
interest. Although in many control systems the variable that we are interested in
minimizing is one of the measured variables, this does not have to be the case.
The controller is defined as the logic or algorithm that is used to guide the actua-
tor. The design of the controller (also referred to as the compensator) is typically
based on a model (mathematical representation) of the system being controlled.
For the application of control design tools, it is sufficient to have an input/output
description of the system, as illustrated schematically in Fig. 1.
Inputs to the model in general include control inputs, disturbances, and sensor
noise, and outputs may include both sensor measurements and predictions of other

System
f y

Fig. 1 Input/output description of a system.


62 D. G. MACMYNOWSKI AND D. R. WILLIAMS

“performance” variables of interest as already described. If there is a single


(scalar) control input and a single sensor measurement, the control system is
referred to as single-input single-output (SISO); if there are multiple inputs and
multiple outputs, the system is referred to as MIMO.
Note that in describing the overall flow control system, the actuators, the
sensors, and the control algorithm can all be reasonably described as parts of
the “control system.” However, for the person designing the control algorithm,
the “system” or plant that is being controlled includes the actuators and sensors,
as shown in Fig. 2. For this purpose, therefore, the model of the system should
describe the response from the control input to the measured output, and should
include any dynamics associated with the actuators and sensors (and also ampli-
fiers, filters, and digital-to-analog converters). It should be noted that the academic
discipline of control theory typically emphasizes the analysis and synthesis of the
control algorithm, and frequently does not include any significant emphasis on the
application-specific actuation or sensing technologies.
Models are typically either in the frequency domain or in the time domain, with
the former only strictly defined for linear systems only and typically useful only
for relatively small numbers of inputs and outputs. The frequency-domain model
is of the form y(s) = G(s)f(s), where f and y are inputs and outputs, s = s + iw is
the (complex) Laplace variable, and G(s) is the transfer function, with G(iw) being
the frequency response of the system. The terms “transfer function” and “fre-
quency response” are often used interchangeably, since an analytic function of a
complex variable is completely described by its response on the iw axis. The
frequency response can be computed from a time-domain model. Frequency-
domain models can also be obtained directly in experiments, by exciting the
system with a known input f, measuring the output y, and computing the ratios of
the Fourier transforms of f and y (with appropriate averaging). System identifica-
tion techniques can also be used to estimate time-domain models from the experi-
mental data signals f and y.
In the time domain, the system can be described by its state variables. These are
a (typically minimal) set of time-dependent variables [which we will denote as a
vector q(t) 僆 Q] that, together with knowledge of all external inputs d to the system
(d including here both disturbances and control inputs), are sufficient to describe
the future time evolution of the system as a first-order differential equation of the
.
form q = f(q, d, t). The equations are then in state-space form. The states include
the relevant flow variables, and may also include any relevant actuator and sensor

System

Fluid
Actuator Sensor
subsystem
f
y

Control

Fig. 2 For the purpose of designing a control algorithm, actuators and sensors
should be considered as parts of the system being controlled.
FLOW CONTROL TERMINOLOGY 63

dynamics, as already noted. Although the state is typically given the symbol x in
the control literature, herein we use the symbol q to avoid confusion with the spatial
coordinate x. As with the use of u as the control input, the use of the symbol x in
the control literature is near universal, and researchers should be aware of our
departure from standard notation. For a linear system, the dynamics can be written
.
as q = Aq + Bd where A and B are matrices. These descriptions are valid for both
time-varying [i.e., A = A(t) and/or B = B(t)] and time-invariant (A and B constant)
systems. A linear time-invariant system (the class for which most control theory is
derived) is abbreviated LTI. For a system described by partial differential equations
(PDEs), the full state is infinite-dimensional. For a typical (Eulerian) discretized
fluid system, the state variables include all of the relevant fluid variables at each
spatial point, which can clearly lead to very high-dimension systems. Model reduc-
tion thus seeks a model with lower state dimension that approximates the higher-
dimensional model with sufficient accuracy. A low-order model can be obtained
either through a model reduction process from a higher-order model (thus yielding
a reduced-order model) or directly from intuition about what relevant physics to
include [examples include Rossiter’s (1964) description of cavity flow physics and
the Moore–Greitzer (1986) model of compressor dynamics].
As with modeling, feedback control design approaches for linear systems can
be grouped into frequency-domain and time-domain approaches. Classical con-
trol refers to frequency-domain design tools that are primarily SISO, roughly
corresponding to tools developed before about 1960 (but still quite appropriate
and useful today). These include graphical analysis tools such as Bode plots,
root locus, and Nyquist or Nichols plots, and design synthesis tools, including
proportional–integral–derivative (PID) and lead and lag compensators. Note that
the majority of control systems in use today are either PI or PID. Modern control
is based on time-domain analysis in state space, and the tools are mostly based on
optimization (hence “optimal” control), with multivariable inputs and outputs
being handled automatically. These tools include quadratic optimal control for
linear systems (LQR, for linear–quadratic regulator; a regulator being synonymous
with a controller) and minimum-variance optimal estimation (Kalman filter). The
combination of these two gives LQG (an acronym for linear–quadratic–Gaussian,
where these are the assumptions required on the plant, the performance metric,
and the statistics respectively). Robustness is a property guaranteeing that essen-
tial functions of the designed system (stability and some level of performance) are
maintained even though the model may not accurately represent the real system
owing to un-modeled dynamics or to changes in the dynamics, or to both. Thus,
robust control design includes any approach that explicitly takes this uncertainty
into account during the design process, and this includes tools such as H• and
mu-synthesis. Many modern control tools are based on minimizing a norm of the
performance, and LQG corresponds to minimizing the H2 norm. (The H2 norm of
a transfer function is the rms of the frequency response.)
The dynamics of many systems have some “slow” time variation, where the
variation is slow with respect to the dynamics being controlled (that is, there is a
separation of timescales) and where the system dynamics relevant to the control
can be reasonably modeled as time-invariant at each operating point. An example
might be where the fluid dynamics of interest varies with the external flow speed.
In this case, the control parameters may be slowly adjusted automatically based
64 D. G. MACMYNOWSKI AND D. R. WILLIAMS

on measurements of the system. If the control parameters are designed offline for
several operating points, and the parameter are interpolated or switched based on
a (slow) system measurement, then the controller is described as scheduled or
gain-scheduled. If the control parameters are updated as a function of dynamic
measurements, then the controller is adaptive. This might take the form of esti-
mating a plant model online with control design based on this model, or might be
model-free in that the algorithm does not depend on having a detailed model of
the plant.
Adaptive controllers are usually nonlinear, and adaptive control typically
implies that the control parameters (e.g., gains) are varied slowly relative to the
bandwidth of the control in order to adapt to slow changes in the dynamics of the
system. Note that different parameterizations of the same controller may lead to
different views on whether the parameters are changing or fixed; in the active
noise control literature, “adaptive feedforward” algorithms for tonal disturbances
can be rewritten as fixed-gain feedback algorithms, as noted in Sec. III. C. A neural
network is one approach for generating a nonlinear adaptive control system
(although adaptation may often be turned off after the “learning” phase).
Of course, there are many other control design approaches not described here.
The performance of a control system clearly cannot be captured by a single
number, but one important number that is useful in describing performance is the
bandwidth. Informally, this is the range of frequencies (minimum to maximum)
over which there is a significant effect. This concept is applicable to describing the
bandwidth of an actuator or of a controller. In many cases, there is only a maxi-
mum frequency (the minimum is zero); in others, there may be a narrow range
of frequencies. For a SISO control system, there are multiple precise definitions
that are not in general identical. If the control objective is command-following,
then the most appropriate definition is the frequency __ at which the response to
a command is at -3 dB (the response is a factor of ÷2 smaller than the command).
This can be measured from the complementary sensitivity transfer function. If the
control objective is disturbance rejection, then the most appropriate definition is
the frequency at which the response
__ to a disturbance is at -3 dB (the disturbance
is attenuated by a factor of ÷2 relative to the response that would have occurred
without control). This can be measured from the sensitivity transfer function. The
sensitivity S and complementary sensitivity T can be obtained from the transfer
function G of the system from input f to output y and the transfer function K of the
controller (from y to f) as S = (1 + GK)-1 and T = 1 - S (see Chapter 5). Often, the
bandwidth is given as the frequency at which the loop transfer function L = GK
has unit magnitude. This is common usage, but is less useful in comparing the
performance of different controllers; this frequency should be referred to as the
loop crossover frequency instead.
Two other terms that are commonly encountered in descriptions of the charac-
teristics of a SISO controller are gain margin and phase margin. These describe
the robustness of the control to uncertainty, and refer to the maximum change in
gain or phase, respectively, that can be tolerated before instability. There is no
simple equivalent for multivariable systems; robustness for these systems is typi-
cally captured through norm-based descriptions.
Several other concepts are relevant in describing the properties of systems to
be controlled, but are probably less critical to those who simply need familiarity
FLOW CONTROL TERMINOLOGY 65

with the control aspects of flow control as opposed to those who need to design
controllers. In control terminology, the system is controllable in finite time if
there exists a control trajectory that can arbitrarily move the state of the system
from any state A to any state B in finite time. The system is approximately con-
trollable if one can get arbitrarily close to state B. Controllability typically cannot
be achieved in fluid mechanics; however, in many cases, it is sufficient if the
system is stabilizable, that is, if all of the unstable modes are controllable (or all
uncontrollable modes are stable) (for a discussion in a fluids context, see Lauga
and Bewley, 2003). Another relevant concept is reachability: states B that can be
reached with some control input are termed reachable. This formalizes the prac-
tical, but not mathematically rigorous, usage of “controllability” that simply
implies the ability to impact the metric of interest. Observability can similarly be
defined both rigorously on the basis of being able to estimate all of the state
variables given the prior time history of the measurement and practically in terms
of the ability to estimate the relevant quantities. The system is detectable if all
unstable modes are observable.
For dynamic systems, stability (in the sense of Lyapunov) is defined as the situ-
ation where small perturbations in the state about equilibrium result in the state
trajectory remaining close to equilibrium [formally, "e > 0, $d > 0 such that
||x(0)|| < d fi ||x(t)|| < e "t > 0, where ||·|| is a norm and the equilibrium point is
assumed to be at x = 0]. This definition can be extended to define the stability of
limit cycles or periodic orbits. Other definitions of stability exist, such as asymp-
totic stability or bounded inputs always resulting in bounded outputs. The fluid
dynamics community makes a further distinction between absolute and convec-
tive instability, where the latter implies that disturbances grow as they convect
downstream. In the control community, the latter case would be viewed as stable
for a finite spatial domain in which the amplitude within the domain remains
finite for any finite initial perturbation. A different non-Eulerian state representa-
tion of a convectively unstable system could be obtained that would be temporally
unstable.

III. Flow Control Definitions


The preceding section described terminology that is generally used in any
control problem. With this brief introduction, we now define a few specific terms
in the context of flow control. Following definitions of flow control and unsteady
flow control, we discuss the concepts of feedback vs feedforward control, closed-
loop vs open-loop control, quasi-steady vs dynamic feedback, and active vs
passive flow control. With these defined, an overall approach to categorizing flow
control systems can be suggested.

A. Flow Control
A possible definition of flow control, from Collis et al. (2004), is: that “Flow
control attempts to alter a natural flow state or development path (transients between
states) into a more desired state (or development path; e.g. smoother, faster tran-
sients).” This definition intentionally puts no constraint on either the spatial scale
of the actuation relative to the flow or the timescale of the actuation relative to the
66 D. G. MACMYNOWSKI AND D. R. WILLIAMS

flow. While it is sufficiently broad to include any concepts that have been given the
label of flow control in the literature, it could also include broader concepts such as
shape optimization that would not typically be considered flow control. It is unlikely
that any definition could be adopted upon which everyone would agree [for
example, the distinction in terminology between shape optimization (not flow
control) and riblets or vortex generators (flow control) is as much historical as
technical]. The subset of flow control that is relevant to this book involves the
integration of tools and ideas from fluid dynamics and control theory.

B. Unsteady Flow Control


Unsteady flow control involves any intentional time-varying effect, such as the
addition of mass, momentum, energy, or vorticity, or shape modulation, including
periodic or quasi-periodic approaches, at a timescale commensurate with the rel-
evant dynamics of the flow. The final distinction on timescale between steady and
unsteady approaches is necessary because any “steady” blowing or suction is not
applied forever. Note that “steady” approaches may have a time-varying effect
due to variation in the system (e.g., aircraft maneuvering impact on vortex genera-
tors); these are still typically considered steady flow control.

C. Feedforward vs Feedback
Some control systems have an actuator command f that depends on some sensor
measurement y from the system; that is, the control action is somehow coordinated
with the system. Control systems that do not depend on any sensor measurement
are a subset of open-loop control approaches, defined in the next subsection. The
dependence of the control system output on a sensor measurement can be either
feedforward or feedback, as sketched in Figs. 3a and 3b, respectively.
Feedforward information does not depend in a significant way on what the
control system has done to the overall system, and is typically intended to give
the controller advance knowledge of disturbances. An example might be where
the control introduces a wave that cancels the disturbance through superposition.
There is feedback if the measurement y changes as a result of the control system
output f; that is, if it depends on f through the dynamics of the plant. This will
always be the case if the measured variable is directly related to the performance
metric that the controller is designed to improve. Aside from the interpretation of
“significant,” these terms are unambiguous. Note that the system in Fig. 3b has a

a) b)
Other Other
disturbances disturbances
System 1
System System

y f f y
Control Control

Fig. 3 Control: a) Feedforward and b) feedback.


FLOW CONTROL TERMINOLOGY 67

feedback loop: f influences the system, which leads to the output y, which changes
f through the controller. There is no such loop in the pure feedforward system
in Fig. 3a.
The definition does not rely on any spatial arrangement between the actuator
and sensor. A control architecture that is intended to be feedforward, with an
upstream sensor giving advance information about the disturbance that will reach
the actuator location, may involve some weak feedback from the actuator com-
mand traveling upstream to influence the sensor response. If ignoring this feedback
path does not significantly alter the system behavior, then the presence of the
feedback loop was negligible, and it is reasonable to describe the system architec-
ture as feedforward.
One control system can involve both feedforward and feedback, as illustrated in
Figs. 4a and 4b. Both representations of the controller use two distinct pieces of
information, y1 and y2. The representation in Fig. 4a, which is common in the active
noise control literature, has been described as “adaptive feedforward”, because the
measurement y1 is thought of as “adapting” the feedforward control law, whereas
the representation in Fig. 4b appears as a controller combining feedforward and
feedback. The topology is equivalent, and the distinction depends only on the algo-
rithms coded in the “control” block (and, indeed, for the algorithms most commonly
used in active noise control, there is no mathematical distinction either).
Note that the term “feedforward” is typically not used if the measurement is not
used at a timescale commensurate with the dynamics. Thus, using an aircraft’s
overall flight speed to modify parameters of a flow control system would typically
be labeled scheduled control rather than feedforward, whereas using an instanta-
neous measurement within the boundary layer upstream of separation could be
feedforward.
“Feedforward” is also used to describe the part of the control command that is
precomputed as a function of the desired reference input for a tracking problem;
in this case, the additional input variable that influences the control response is not
a sensor measurement, but a desired reference input.

D. Open-Loop vs Closed-Loop
The strictest definition of a closed-loop system is that it has a feedback loop. A
control system is thus closed-loop if a sensor is used to change the output of an
actuator, which in turn affects the sensor through the dynamics of the plant. If no
such closed loop exists, the system is open-loop (e.g., if the actuator does not

a) b)
System 1
System System 1
System

y2 f f
Control y1
y1 Control
y2

Fig. 4 Control system: a) “Adaptive feedforward” control and b) feedforward plus feed-
back control.
68 D. G. MACMYNOWSKI AND D. R. WILLIAMS

affect the sensor in any significant way). Based on this definition, open-loop flow
control strategies include 1) those that are predetermined, with no sensor mea-
surement at all (e.g., a fixed sinusoidal excitation of fixed amplitude and fre-
quency), 2) those that are purely scheduled (e.g., a sinusoidal excitation whose
frequency and amplitude are chosen as a function of aircraft flight speed), and
3) purely feedforward systems (e.g., where a cancelling signal is introduced based
on an upstream measurement of the disturbance to be cancelled).
The term “closed-loop” has also been used in the literature to describe any
control system that depends on a system measurement regardless of whether it is
feedforward or feedback, and thus “open-loop” would refer only to those systems
that do not depend on any sensor measurement. However, this usage will lead to
confusion, because the term is unambiguous to most people in the control com-
munity, and therefore the strict definition should be followed. Figure 5 illustrates
the topological options for a control system; the options in Figs. 5a and 5c are
clearly open- and closed-loop, respectively, whereas that is Fig. 5b, is considered
by some to be closed-loop but should be described as open-loop.
In receding-horizon model-predictive control, optimization is used to compute
the future control trajectory based on the current sensor measurement, this is
applied for some period of time, and then a new trajectory is computed based on
new sensor measurements. This is therefore a closed-loop control strategy (since
the control depends on measurements that in turn depend on the previously applied
control), although it has been occasionally and incorrectly described as being
open-loop because the optimization step itself computes an open-loop trajectory.
An additional potential source of confusion results from the common usage of
the term “open-loop” in the control community to refer to the uncontrolled system
(this presents less of a communication problem in applications where open-loop
control strategies are not common). Consider the system shown in Fig. 6, where
the control loop has been broken, and the actuator command f could be zero
or could be separately determined for the purposes of measuring the dynamics of
the system. The system with zero control (with the response y being due only to the
exogenous input d) should be referred to as uncontrolled rather than “open-loop”,
to avoid confusion. Note that this might not be the same as the baseline flow
response without actuators or flow control devices installed. For example, a syn-
thetic jet may still act as a Helmholtz resonator or boundary-layer trip when
powered off. The loop transfer function (for a linear system) from some specified
(but open-loop!) applied actuator command f to the actuator command w that is
computed from the control law is occasionally referred to by some in the control

a) b) c)
System System 1 System System
f
y f y
f
Control Control Control

Fig. 5 Topological options for a control system: a) open-loop control, b) coordinated


(feedforward or scheduled), and c) closed-loop control.
FLOW CONTROL TERMINOLOGY 69

d
System
f
y

w Control

Fig. 6 System with broken control loop.

community as the open-loop transfer function. It is more common to refer to the


response from f to y as the open-loop response. Since people are likely to continue
to refer to any transfer function from any input on Fig. 6 to any output as an “open-
loop” transfer function, it is essential to be specific in order to avoid confusion.
As already noted, for linear systems, for which much of control theory has been
developed, the dynamics of the system cannot be changed with open-loop control.
However, many flow control problems are interesting precisely because this is not
true for nonlinear systems. Examples include unsteady open-loop forcing to
modify separation dynamics, and cavity flow dynamics.

E. Quasi-Steady and Dynamic Feedback


Flow control closed-loop systems can be loosely categorized on the basis of
timescale. If the control commands (e.g., the amplitude and phase of a peri-
odic excitation) are adjusted slowly compared with the dynamics of the flow, then
the control system could be referred to as a quasi-steady controller, whereas if
the control parameters are adjusted on a timescale commensurate with that of the
dynamics, then the control system could be termed dynamic (Cattafesta et al.,
2003). However, this is not standard terminology in control, as both are dynamic
control systems. The distinction depends of course on the dynamics of interest,
and thus the ambiguity results from the choice of the system or plant being con-
trolled. Choosing the forcing amplitude of flow control actuation to control the
motion of an aircraft might be “dynamic” relative to the aircraft dynamics, but
“quasi-steady” relative to the separation dynamics. However, from a flow control
perspective, it is the latter phenomena that are of interest. Thus, a more accurate
characterization might be to distinguish between low- and high-bandwidth con-
trol, relative to the relevant frequencies of the dynamics. The distinction between
these cases is clearly relevant, although a single adjective cannot rigorously clarify
it. The distinction is also not judgmental; low-bandwidth control is clearly valu-
able, for example in tuning control parameters using extremum-seeking adaptive
algorithms (Banaszuk et al., 2004).

F. Active vs Passive Flow Control


Flow control strategies are frequently categorized as active or passive. These
terms do not have any clearly accepted rigorous definitions, but nonetheless are
commonly used. Typically, the classification is based on energy addition, on
whether the control is steady or unsteady, or on whether there are parameters (e.g.,
70 D. G. MACMYNOWSKI AND D. R. WILLIAMS

an oscillation frequency) that can be modified after the system is built. We recom-
mend energy addition as the determination of whether a given approach is termed
active or passive.
In making a distinction based on energy addition, active flow control can be
either steady or unsteady, but requires external energy (electrical or mechanical),
whereas passive flow control, such as the use of vortex generators or riblets, does
not require external energy. This distinction, however, is inherently ambiguous,
since the term “external” depends on the physical and temporal boundaries of the
subsystem being considered, and thus some judgment and pragmatism should
be applied.
Consider the cartoon examples in Figs. 7a–7c. A synthetic jet connected to
external power is “active flow control”. If one used a tiny turbine to power a genera-
tor to drive one’s flow control system, it might be self-contained, but should prob-
ably also be considered active. If one bypassed the conversion between fluid and
electrical power, for example by using fluidics to convert mean energy into unsteady
energy, then the system might be conceptually identical, but has often been labeled
passive. Note that feedback control systems can and have been implemented using
only fluidics, and thus such a system has the same potential to destabilize the flow
as a similar system using electrical power. If external energy is required, but only
infrequently, such as for “pop-up” mechanical vortex generators, then it is probably
more useful to describe the flow control as passive, since that more accurately
describes the mechanism of control.
An alternate but uncommonly used definition would include all steady flow
control as passive, regardless of whether external energy is supplied (e.g., by
blowing or suction), and all unsteady flow control as active. Some flow control
devices (e.g., riblets and fixed vortex generators) are passive regardless of the defi-
nition, others (synthetic jets) are active regardless of the definition, and others will
always be in a gray area.
One particular program on controlling compressor stall introduced a finer
delineation, with “level 0 control” being purely passive, “level 1” involving open-
loop scheduling, “level 2” involving quasi-steady low-bandwidth feedback, and
“level 3” involving high-bandwidth dynamic feedback.

G. Flow Control Classification


A possible overall categorization of flow control approaches, as described
through the last several definitions, is shown in Fig. 8.

a) b) c)

Fig. 7 Flow control: a) “Active” synthetic jet, b) flow-powered synthetic jet, and c) fluidic
unsteady jet.
FLOW CONTROL TERMINOLOGY 71

Fig. 8 Categorization of approaches to flow control.

There is often a perceptual distinction in terms of the implied complexity,


robustness, and how advanced the flow control system is. Although these percep-
tions clearly influence decisions on risk and on funding, it is more appropriate to
move beyond these simple classifications and consider each system on its own
merits, rather than on the basis of a few ill-defined adjectives.

H. Flow Control “Boundaries”


The definitions given in this chapter are useful in communicating and compar-
ing flow control ideas. It is also important to understand what to include in a flow
control system, and what is not flow control.
The first definition given in this chapter, in Sec. II, noted the importance of
establishing the spatial and temporal boundaries of a flow control system, and this
issue has arisen several times in the preceding discussion of the classification of
systems. The most important aspect of this issue is not the terminology used, but
the need to be honest about the overall benefit—that is, to include the penalties
associated with the energy input required, in addition to the benefits obtained. If the
energy input required to remove a flow separation is larger than the loss associated
with the separation, then the flow control system is probably not beneficial.
One final note is in order regarding on what should not be included as flow
control. A glance through the table of contents of this book and at the history of
flow control in Chapter 1 gives an excellent set of examples of what is flow con-
trol. Many other engineering applications involve controlling the motion of flow,
but are more appropriately categorized as flow management.
Chapter 4

Role of Instability Theory in Flow Control

Vassilis Theofilis*
Technical University of Madrid, Madrid, Spain

I. Introduction
It is to be expected that a chapter on the role of instability theory in flow control
will dwell on the interconnection between two of the broadest concepts in fluid
mechanics. Indeed, as testified by the other chapter in this volume, flow control is
an extensive and ever-expanding area of research and application. From the outset,
is important to be aware of two possible misconceptions. First, one might think
that there is now a complete understanding of fluid flow instability, since vast
amounts of information have been amassed by consistently high levels of effort
over more than a century (Criminale et al., 2003; Drazin and Reid, 1981; Schmid
and Henningson, 2001). Second, it has been claimed that the field of fluid
mechanics has reached maturity, a corollary being that the expert community
should now be capable of applying existing knowledge in order to modify a certain
flow in a desired manner.
This chapter is intended primarily to dispel both of these myths and demonstrate
that knowledge of flow instability physics, or at least exploitation of an instability
mechanism (which may not necessarily be understood), is indeed indispensable to
devising successful and robust flow control methodologies; moreover, the effi-
ciency of a flow control methodology based on flow instabilities may be maximized
by using receptivity information to minimize actuator requirements. Information
on the instabilities supported by a flow system is contained in the equations of
motion, so emphasis will be put on the properties of the operator describing the
linearized Navier–Stokes equations (LNSE). It will be stressed that the concept of
flow instabilities must be put in its correct context: the scope of classic (modal)
linear theory (Drazin and Reid, 1981) which deals with the identification and analy-
sis of the eigenspectrum, must be broadened in order to include large areas of flow
instability that until recently have gone unnoticed. Concretely, the concepts of

Copyright © 2008 by the author. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*School of Aeronautics; vassilis@aero.upm.es.

73
74 V. THEOFILIS

transient growth and adjoint-based analysis and optimization will be highlighted;


both areas have flourished over the last decade—mainly using one-dimensional
background flows as testbeds—and have provided a theoretical basis for dealing
with the questions that modal linear instability theory has left unanswered. Further,
the relatively recent nonparallel flow instability theory, comprising both modal and
nonmodal aspects, will be discussed, because this is an area of prime relevance to
the complex configurations encountered in applications.
Turning to flow control, at a macroscopic level, this area may be seen as an
optimization problem involving flow around an object (e.g., a full configuration).
A well established theoretical approach exists for this problem and is depicted in
Fig. 1. It involves the solution of two classes of problems along the two distinct
direct (left column/downward arrows) and adjoint (right column/upward arrows)
paths shown in this figure, at three possible levels of complexity, shown as three
rows of horizontal boxes in Fig. 1. At the top level, the methodology includes
aerodynamic shape optimization (Giles and Pierce, 2001b; Jameson, 1988, 1989;
Lions, 1971), indicated by dotted lines, in which full-configuration performance
requirements determine the cost function defined within the optimization algo-
rithm. The full optimization procedure, in which flow control is based on control
of flow instabilities (Pralits and Hanifi, 2003), includes two further steps. First,
solutions of the appropriate equations of motion are obtained in parts of the flow
where the introduction of viscosity is essential to describe the phenomena observed,
for example, in areas of flow separation. The relevant solutions can be obtained from
(similar/nonsimilar, interactive/non interactive, as appropriate) solutions to the

Fig. 1 Schematic representation of the overall optimization algorithm for complex


configurations.
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 75

boundary-layer (BL) equations in regions of the flow where this approximation is


permissible (e.g., fuselage/wing surface), or from solution to the full Navier–Stokes
equation in all other parts of the flowfield. This step provides the background flow,
which may be steady- or time-periodic, laminar or time-averaged turbulent flow,
respectively referred to as basic or mean flow. The background flow thus gener-
ated may be analyzed in the final step of the full optimization algorithm where
direct and adjoint LNSE are employed. An obvious (but pedestrian) optimization
approach using only direct LNSE would require the computation (by solution of
the full nonlinear Navier–Stokes equations) of the basic/background flow field
corresponding to each modification of the flow configuration that is deemed inter-
esting/appropriate; subsequently, local or global instability analysis of each
background field would be performed for a range of parameters and the results of
the amplification rates monitored in order to assess the success/failure of the
introduced flow modification. Though cumbersome and expensive, especially in
complex flow problems, this direct approach has been successfully applied in the
context of global instability analysis (Morzynski and Afanasiev, 1996) to the
archetypal problem of the control of the wake of the circular cylinder by a
Strykowski wire (Strykowski and Sreenivasan, 1990).
In contrast, and in a manner analogous to aerodynamic shape optimization
(Giannakoglou, 2004), the main attraction of the adjoint-based theoretical approach
to flow control is that it removes the need to perform an instability analysis of every
background flow corresponding to a modification introduced in the flow. Adjoint-
based flow control treats the flow modification as a feedback controller which intro-
duces linear perturbations to the leading (direct) eigenmodes. Extensive discussion of
the various aspects of flow control via control of flow instabilities may be found in
other chapters of this volume (Bewley, 2001; Collis et al., 2004; Kim and Bewley,
2007). A full discussion of linear control theory may be found in other chapters of
this volume; here, elements of linear control theory are included for completeness,
especially to make the link with linear instability theory on the basis of the equations
solved by the respective problems. With a view to applicability to real configurations,
instability analysis in the present context should be understood as that of isolated
regions in the flow, control of which may improve overall performance. The focus of
this presentation is on a parameter-free theoretically founded methodology to what is
now known as active flow control, via a combination of solutions to the direct and the
adjoint eigenvalue problems; to the author’s knowledge the first work which has
elucidated the interconnection between linear instability and flow control in complex
flows is the work of Hill (1992) which successfully addressed the longstanding prob-
lem of control of cylinder wake. Subsequent work by the same author (Hill, 1995)
has applied the adjoint-based methodology to control instabilities in the flat-plate
boundary layer. In a parallel development, Joslin et al. (1995b, 1997) have employed
direct simulation of the full equations of motion to the latter problem; a complete
account of the rapidly-growing field of active flow control via direct numerical simu-
lation may be found elsewhere (Joslin, 2001). The interested reader is also referred to
recent works by Høpffner (2006), Chevalier et al. (2007) and references therein for
current efforts in the area of non-parallel flow control.
It is the intention of the present chapter to keep the discussion at a level of
abstraction which permits incorporation of the diverse aspects of flow instability
and control within a unique framework. In this respect, what is commonly known
76 V. THEOFILIS

as a passive control approach (e.g., boundary-layer control by introduction of


roughness) would be understood as a steady modification of the background flow,
for which an instability problem would be solved and its results compared with
those of the unmodified baseline background flow configuration. On the other
hand, theoretical active control (e.g., of Tollmien–Schlichting waves) requires the
introduction of the concepts of direct/adjoint optimization and feedback control.
One would also not differentiate from a theoretical point of view between control
of a flow in which clear mode separation exists from one in which a wavepacket
must be considered, although the level of theoretical and numerical complexity of
the two problems may well be different. It should be stressed that, whereas until
recently instability analysis and flow control was mainly confined to solutions of
problems of primarily academic interest, recent advances in both instability theory
(Schmid and Henningson, 2001) and numerical algorithms for the solution of
large eigenvalue problems (Theofilis, 2003) permit addressing previously intract-
able configurations of engineering significance in or around three-dimensional,
spanwise periodic objects of arbitrary two-dimensional cross-section, such as
airfoils, cavities and back-steps, chevron nozzles, and trailing-vortex systems.
In what follows, the theoretical tools currently employed in linear instability
analysis and linear flow control are exposed in a summary fashion. From the
outset a distinction must be made between two classes of theoretical flow control
approaches currently employed: on the one hand adjoint-based optimization
including base flow modifications, receptivity and optimal control, and on the other
hand linear control theory including LQG, H2, primarily based on solution of the
Riccati equation. The present chapter will deal mostly with the first approach,
whereas the next chapter will expose the second in some detail. Regarding instabil-
ity theory, the classic modal analysis concept is embedded into the more general
stability theory applicable to both modal and non-modal perturbations. The discus-
sion covers temporal and spatial linear stability problems alike, but the dimension-
ality and properties of the linear operator governing the different linear stability
problems are highlighted. Section II deals with linear modal and non-modal insta-
bility theory, while Sec. III introduces one of the two theoretical alternatives to flow
control: receptivity and adjoint-based optimization. A section on linear-systems
control in the next chapter closes this topic. Flow control applications related to
the control of flow instabilities are exposed in a selective manner in Sec. IV. An
assessment of the current situation as well as a discussion on expected future
directions in this rapidly expanding field, offered in Sec. V, close the present
chapter. Apologies are offered to the authors of a large body of interesting work
which has not been cited, either because their work belongs to the same class as
that of a cited paradigm, or because this chapter is written with a clear theoretical
bias as a complement to more applied chapters in the present volume.

II. Elements of Instability Theory


The control of flow instabilities is based on the exploitation of small-amplitude
perturbations; such perturbations can be described by the linearized Navier–Stokes
equations. Monitoring the linearized part of the equations of motion is mandatory
since, for shear flows, the nonlinear terms in the equations only serve to redistribute
energy amongst different spatial scales (Joseph, 1976; Schmid, 2000) and are not
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 77

directly related to growth or decay of perturbation energy. Linearization of the


equations of motion may be performed around some background flow; the term
“background” flow is introduced in order to describe collectively steady or
unsteady laminar basic flow or ensemble- or time-averaged turbulent flow. Basic
flows may be provided either analytically or through highly accurate one-, two-,
or three-dimensional direct numerical solutions of the equations of motion, with-
out turbulence model, exploiting temporal or spatial invariances in the flow. On
the other hand, the mean of a turbulent flow may be obtained either by experiment
or from long-time integrations of three-dimensional numerical solutions of the
full equations of motion without (cf. DNS) or in the presence of (cf. LES, DES,
RANS) some form of turbulence model.

A. The Initial-Value Problem and its Formal Solution


In using the term “small-amplitude perturbation” it is stressed at this point that
we refer to solutions of the initial-value problem
du = Au
___ (1)
dt
where u(t) is the state vector comprising the amplitude functions of the linear
perturbations. The linearized operator A depends on the background state; the
latter may be independent of time, in which case the autonomous system (1) has
the explicit solution

u(t) = eAtu(0) (2)

where the matrix exponential is known as the propagator operator. In case the
background flow exhibits an arbitrary dependence on time (Farrell and Ioannou,
1996), there still exists a propagator matrix, F[t,0], which solves Eq. (1):

u(t) = F[t,0]u(0) (3)

In the particular case of a time-periodic background flow, for which Floquet theory
is applicable (Herbert, 1984; Barkley and Henderson, 1996), this matrix is denoted
the monodromy matrix (Karniadakis and Sherwin, 2005).
Concentrating on the autonomous system first, one further step may be taken by
applying the spectral theorem; the matrix discretizing the linearized operator A
may be diagonalized if and only if the matrix is normal. A normal operator is one
which commutes with its adjoint* and is completely described by its eigenvalue
system. In the case of a normal operator

A = ULU-1 (4)

U being a unitary matrix, i.e., one for which UU* = U*U = I holds, U* being the
adjoint of U and I being the identity matrix. The diagonal matrix L contains the

*See classic textbooks (Morse and Feshbach, 1953) regarding the definition of the adjoint of a
linear operator.
78 V. THEOFILIS

eigenvalues of A, whereas the eigenvectors of A, contained in U, are orthogonal.


The solution of the autonomous system may then be expressed as

u(t ) = UetL U -1u(0) (5)

and the orthogonality of the eigenvectors ensures that the state of the flow at time
t may be fully determined by the initial condition, an eigenvalue, and the corre-
sponding eigenvector; one refers to modal growth in this situation.
An alternative description, valid for both normal and non-normal matrices A,
isolates three situations, two described by the state at which the flow is found in
the limits t Æ 0 and t Æ •, and a third intermediate situation at finite times.
Although only the limit t Æ • is related to description (4), all three flow states
may be monitored by considering the growth s, in times of an arbitrary initial
perturbation u(0), through
*
(e A te At u(0), u(0))
s2 = (6)
(u(0), u(0))

Implicit here is the definition of an inner product, (·, ·), and the associated
adjoint, A* of the linearized operator A. Neither of these is uniquely defined and
one must appeal to the physical interpretation of the linearized perturbations in
order to attach significance to the choice made. The discussion is completed by
introducing the singular value decomposition (SVD) of the propagator operator

eAt = USV* (7)

Here the unitary matrices V and U respectively comprise (as their column
vectors) initial and final states, as transformed by the action of the propagator
operator, whereas S is diagonal and contains the associated growth s as the cor-
responding singular value. Both modal- and non-modal growth may be studied
by this alternative description.
The matrix exponential function first appearing in Eq. (2) is key to instability
analysis and will also be encountered in linear flow control. Reliable results in
both areas require numerical methods for the accurate calculation of this function,
itself a matrix of the same dimension as the linearized operator A. Several meth-
ods have been proposed in the literature for the computation of the matrix expo-
nential function; the two articles by Moler and Van Loan (2003) with the same
title, 25 years apart, are essential reading in this context. The simplest method
requires knowledge of the eigenspectrum of A but fails in cases where the eigen-
vectors are near parallel, as is often the case in shear-flow instabilities. Alternatives,
based on truncated power series, of Padé approximations are viable when the
leading dimension of A is numerically manageable. In that case direct methods
for the evaluation of the matrix exponential function are feasible, otherwise
(Krylov subspace) iteration must be employed. Expanding somewhat on this
point, it should be noted that the propagator operator itself can actually represent
a Navier–Stokes solver. The associated so-called matrix-free or time-stepper
methods (Karniadakis and Sherwin, 2005) are presently being successfully
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 79

employed to both the instability analysis and the control problems. Note that the
properties of the matrices involved are crucial in determining the accuracy of
the solutions obtained. This discussion naturally leads to a closer inspection of the
linear operator, to which we turn next.

B. Dimensionality of the Linear Operator A


The dimensionality of the background flow determines that of the linearized
operator A. A background flow may comprise one, two, or all three velocity-
components—alongside pressure—and, if compressible, density/temperature
variations. Potential for confusion arises when reference is made in the literature
to control of two- or three-dimensional flows/perturbations, irrespective of the
number of spatial directions considered as homogeneous in the underlying back-
ground flow. The classification followed here for A is a result of the dimensional-
ity of the background flow; it is referred to as an N-dimensional problem if
M = 3 - N homogeneous spatial directions exist in the background flow, i.e., M
directions are treated by a Fourier expansion, while N spatial directions are fully
resolved (in a coupled manner if N ≥ 2). The same nomenclature convention is
adopted for the amplitude functions of small-amplitude perturbations which
inherit the dimensionality of the background flow.
The full picture is presented in Table 1 which, for the sake of simplicity, refers
to the autonomous system and to laminar (incompressible or compressible) flow.
It is further referred to as a quasi N - d problem if the dependence of perturba-
tions on the Nth spatial direction is much weaker (in an asymptotic sense, using
the Reynolds number as an expansion parameter) than that on the other N - 1
spatial directions; the latter are fully resolved. Concretely, natural convection
problems which may develop in the absence of background flow are referred to
as 0 - d problems, while instability analysis and control of disturbances develop-
ing upon a one-dimensional basic flow (a “profile” such as encountered in models
of the flat-plate boundary layer or isolated vortices) is referred to as a 1 - d problem.
This refinement in the terminology utilized is important since, on the one hand,
the vast majority of analyses in the literature have focused on the simplest (0 - or
1 - d) background flows, which result in analytically and/or numerically tractable
problems, while most realistic applications correspond to at least quasi 2 - d
problems.
When a new flow control application is encountered, the first consideration
would be identification of the background state among the different classes
shown in Table 1. The most efficient numerical path for the numerical solution of
the associated theoretical flow control problems follows immediately from this
classification.

C. Operators A and A* in Incompressible Flow


For convenience in the presentation, especially of the (continuous) adjoint
problem, the incompressible equations of motion are discussed. The equations are
linearized around a steady laminar basic state (ū, p̄)T

u = ū + eû; p = p̄ + ep̂ (8)


80
Table 1 Decompositions pertinent to the direct linearized equations of motion

Eigenmode
Amplitude
Theory Basic state function Phase function

TriGlobal q̄(x1, x2, x3) + q̂(x1, x2, x3) · exp Q3d(t) ; Q3d = -Wt

V. THEOFILIS
x≤
Parabolized stability equations-3D q̄(x≤1, x2, x3) + q̂(x≤1, x2, x3) · exp Q≤2d(x≤1; t) ; Q≤2d =x=xÚ a(x) dx - Wt
• 0

BiGlobal secondary theory q̄(x1, x2, x3¢; t¢) + S


n=-•
q̂(x1, x2, x3¢, t¢) · exp Q¢2d(t¢) ; Q¢2d = st
BiGlobal q̄(x1, x2) + q̂(x1, x2) · exp Q2d(x3; t) ; Q2d = bx3 - Wt
x≤

Parabolized stability equations q̄(x≤1, x2) + q̂(x≤1, x2) · exp Q≤1d((x≤1, x3; t) ; Q≤1d =x =xÚ a (x) dx + bx3 - Wt
• 0

Herbert secondary theory q̄(x1¢, x2; t¢) + S


n=-•
q̂(x1¢, x2, t¢) · exp Q¢1d(x3; t¢) ; Q¢1d = bx3 - st
Rayleigh/Orr–Sommerfeld q̄(x2) + q̂(x2) · exp Q1d(x1, x3; t) ; Q1d = ax1 + bx3 - Wt
Internal gravity waves/Rayleigh– 0 + q̂(x2) · exp Q1d(x1, x3; t) ; Q1d = ax1 + bx3 - Wt
Taylor/Benard instabilities
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 81

with e 
 1. The LNSE of motion

∂uˆ
+ Auˆ + —pˆ = 0 (9)
∂t
— · uˆ = 0 (10)

are obtained for the determination of the small-amplitude perturbations (û, p̂)T.
Using Einstein summation and permitting inhomogeneous three-dimensional
background and perturbation fields without special symmetries, the components
of the (direct) linear operator are

∂uˆi ∂u 1 ∂2uˆi
Ai uˆ j = u j + uˆ j i - (11)
∂x j ∂x j Re ∂x 2j

A Euler–Lagrange identity is constructed from first principles and the definition


of an inner product. It reads (Morse and Feshbach, 1953; Hill, 1992):

ÈÊ ∂uˆ ˆ * ˘ È Ê ∂u * * * ˆ ˘
ÍÁË ∂t + Auˆ + —pˆ ˜¯ · u + — · uˆ p ˙ + Íuˆ · ÁË ∂t + A u + —p ˜¯ + pˆ — · u ˙
* * *

Î ˚ Î ˚
∂ (12)
= (uˆ · u* ) + — · P(uˆ , u* )
∂t

The Euler–Lagrange identity serves to define both the linearized adjoint opera-
tor A* and the adjoint boundary conditions, the latter via the bi-linear concomitant,
P(q̂, q*). Assuming incompressibility, the linearized equations of motion for the
determination of the small-amplitude adjoint perturbations (u*, p*)T in the general
three-dimensional case are readily identified in Eq. (12) as

∂u *
+ A* u* + —p* = 0 (13)
∂t
— · u* = 0 (14)

Explicitly, the components of the adjoint linear operator and the the bi-linear
concomitant are respectively

∂u *i ∂u j 1 ∂ 2u *i
Ai*u *j = u j - u *j + (15)
∂x j ∂xi Re ∂x 2j

and

1 Ê ∂u*j ∂uˆ ˆ
Pi = ui (uˆ j u*j ) + Á uˆ j - u*j j ˜ + uˆi p* + pu
ˆ i* (16)
Re Ë ∂xi ∂xi ¯
82 V. THEOFILIS

A second inhomogeneous adjoint problem has been defined by Hill (1992) to


determine the modifications that the amplification rate experiences on account of
small modifications to the background state; some details will be exposed in the
following sections, while the interested reader is referred to the original work for
a full discussion. More recent references in the same spirit include the work of
Giannetti and Luchini (2007), the recent reviews of Kim and Bewley (2007) and
Schmid and Henningson’s (2001) book.

D. On Instability Theory and Turbulent Flows


The relevance of linear stability theory ideas to turbulent parallel free-shear-
flow predictions has been demonstrated, among others, in the celebrated work of
Gaster, Kit and Wygnanski (1985). Given the inflectional nature of the one-di-
mensional mixing-layer profile studied by these authors, the simple inviscid
Rayleigh stability equation has been utilized as a theoretical basis, and excellent
comparisons between theoretical predictions based on this model and experi-
mental results have been obtained. On the other hand, in the case of turbulent
parallel wall-bounded shear flows (again dependent upon one spatial coordinate
alone), theoretical tools for the stability analysis have been developed by Hussain
and Reynolds (Hussain and Reynolds, 1970, 1972; Reynolds and Hussain, 1972).
These authors have introduced the concept of coherent structures and suggested
that any flow quantity, f(x, t), be decomposed into a time-averaged (mean), f¯, a
coherent, f˜, and incoherent, f ¢, parts of the turbulent flow, all three schematically
depicted in Fig. 2.*

Fig. 2 The triple decomposition concept, schematically depicting the mean f¯, a
coherent f˜, and incoherent f´, parts of the turbulent flow (Hussain and Reynolds,
1970, 1972; Reynolds and Hussain, 1972).

*The symbols f¯ and f˜ are used here for consistency with the original references and should not be
confused with their meaning throughout the chapter for basic flow and amplitude function of perturba-
tions, respectively.
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 83

The first part, f̄, may be calculated from measurements (or simulations) which
also provide models for the third part, f¢, of the decomposition. The stability analysis
provides predictions for the spatial structure and frequencies of the coherent struc-
tures, f̃. Concretely, substitution of the triple-decomposition into the equations of
motion results in a nonlinear system of equations that needs to be closed. Unlike
classic turbulent closure, where the flow is decomposed into mean and turbulent
parts alone, the appearance of the coherent structures requires a model for their
description. Borrowing ideas of classic instability theory of laminar parallel flows,
Hussain and Reynolds applied the Orr–Sommerfeld equation to determine the
coherent structures after having modelled the small-scale turbulent flowfield using
an eddy-viscosity concept. In doing so, an unknown function of the (resolved, in
their plane channel configuration) wall-normal coordinate has been introduced and
has been determined from the turbulent channel data.
In order to apply these ideas to non-parallel flows of engineering significance,
as appearing in/over complex geometries, an extension of the triple-decomposition
concept along the lines of BiGlobal theory is necessary; to date no such effort has
been undertaken.

III. Multiple Role of Adjoints in Flow Control


Two major lines of distinct but related approaches to flow control via control of
flow instabilities are currently in use. One is based on identification and exploita-
tion of flow sensitivities via receptivity analysis and adjoint-based optimization
and the second is a linear-systems approach to flow control; both are briefly intro-
duced here.

A. Receptivity
Up to this point it has been seen that, in the case of normal matrices, the eigen-
value decomposition produces unitary matrices whose eigenvectors are mutually
orthogonal, while in the case of non-normal matrices, consideration of a direct
and an associated adjoint eigenvalue problem is necessary. By construction, the
sets of direct and adjoint eigenvectors are mutually orthogonal and, in order to
distinguish from the case of normal matrices, are referred to as the bi-orthogonal
set (Tumin, 2003; Tumin and Fedorov, 1984).
While Eq. (7) may be used in order to predict the behavior of linear perturba-
tions at finite times, the bi-orthogonal set of eigenvectors may be used to identify
sensitivities in the flow by solving the receptivity problem. What is meant by
“sensitivities” is locations in the field where excitation (control) may have maxi-
mum effect. The receptivity problem is solved by augmenting the LNSE system
Eqs. (9–10) by momentum sources, mass and boundary condition variations. It
should be noted that this is a particular case of augmenting the equations which
could, in principle, be extended to include an arbitrary unsteady forcing. If the
direct and adjoint linear perturbations are taken to exhibit a harmonic dependence
in time,

uˆ = u exp(iw t ), u* = u ¢ exp(-iw t ) (17)


84 V. THEOFILIS

the same assumption must be made of the forcing terms. The “forcing” problem
then reads
-iw uˆ + Auˆ + —pˆ = fˆ (18)

— · uˆ = fˆ (19)

alongside appropriate boundary conditions, such as no-slip conditions for walls


and free-stream conditions. Here f̂ and f̂ respectively represent the amplitude
functions of harmonic mass and momentum injection. The (three-dimensional, in
general) amplitude functions of the direct problem are expanded into a sum of M
discrete eigenmodes
M
u = Â am u m + s (20)
m=1

where s̃ is used to include potential contributions from the continuous spectrum


and this expansion is substituted into a modified Euler–Lagrange identity which
takes into account the forcing terms. Conservation of the bi-linear concomitant
and integration along a spatial direction where homogeneous boundary condi-
tions may be imposed leads to the explicit form of the coefficients am that govern
the receptivity problem. The latter may be completely determined using on the
one hand information on the disturbance source and on the other hand the bi-
orthogonal set of solutions of the direct and adjoint homogeneous problems. As
such, a key feature of a particular solution to the adjoint system is the determi-
nation (“filtering”) of the amplitude of the corresponding direct mode. This
approach was first employed in the context of a flow instability problem by
Tumin and Fedorov (1984) and will be discussed further. Note also that the
formulation so far makes no reference to the dimensionality of the operators and
is equally applicable to all problems described in Table 1. In what follows, refer-
ence to concrete examples stemming from presently available analyses which
involve up to two-dimensional basic states will elucidate the concepts discussed
in this section.

B. Adjoint-Based Optimization
At this point concepts from adjoint-based optimization are introduced. Two
branches of this field are of interest in an instability control context, namely con-
trol of linear ordinary-differential equations (ODEs) and linear partial-differential
equations (PDEs). Control of ODEs is relevant to instabilities developing upon
one-dimensional (parallel) basic flows, which is where most of the work to date
has been performed. Control of PDEs is applicable to instabilities of both weakly-
non-parallel (governed by equations of the PSE type) and essentially non-parallel
flows (governed by Bi-Global or Tri-Global analysis) and is a rapidly growing
area, being most relevant to applications.
The first key idea at the heart of an adjoint-based optimization scheme is the
definition of a functional; again, the well studied shape-optimization field serves
to contrast ideas. Whereas in this area one aims at minimizing drag and translates
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 85

this requirement into minimization of pressure coefficient through specification


of a functional involving a target pressure, in control of instabilities it is typi-
cally the kinetic energy of perturbations that is defined in the functional. The
linear system to be analyzed and its associated adjoint linear system have the
generic form
Aû = f, A* u* = g (21)

Both systems are defined over an arbitrary domain W and are subject to homo-
geneous (respectively, direct, and adjoint) boundary conditions for which f and g
are chosen so as to ensure well-posedness. An integral inner-product, denoted by
(·, ·), is defined over W; it is used to define the functional (scalar objective function),
J, through
J = (u*, f ) = (u*, Aû) = (A*u*, û) = (g, û) (22)

At this point it is essential to note that the functional is calculated through


inner-product computations between the solution vector, u*, of the adjoint problem
and the right-hand-side vector f associated with the direct problem. The efficiency
of the adjoint-based optimization technique derives from the fact that there exist
multiple vectors f, each corresponding to different design parameter values in the
direct problem, while there only exists a single objective function, g. Each itera-
tion in a design cycle requires a single adjoint calculation [solution of the adjoint
linear system in Eq. (21)], as well as several functional evaluations [inner-product
computations in Eq. (22)], as opposed to a large number of direct calculations
[solutions of the direct linear system in Eq. (21)] equal to the number of design
parameters.
The second key idea in adjoint-based optimization is the computation of the
gradient of the functional —a J, a denoting a set of design variables. The flow
solution, q̂ = (û, p̂), and design variables a satisfy the linearized equations of
motion and appropriate boundary conditions, written as a constraint

L(q̂, a) = 0 (23)

A typical design variable a in the context of flow control by control of flow


instability (Hill, 1992) is the global amplification rate w resulting from intro-
duction of a Fourier decomposition in time in the BiGlobal instability problem
discussed earlier. Linearization of the equations of motion around a basic flow q̄
results in

dJ ∂J dqˆ ∂J
= + (24)
da ∂qˆ da ∂a

The flow sensitivity dq̂ /da satisfies the linearized constraint

∂L dqˆ ∂L
+ =0 (25)
∂qˆ da ∂a
86 V. THEOFILIS

which results in an expression for the evaluation of the sensitivity of the functional
to perturbations in the design variables which has the same form as Eq. (24) sub-
ject to the linearized equations

Aq̂ = f (26)
with A  ∂L/∂q̂ and f = -∂L/∂a.
An equivalent (Lagrangian) description augments the functional through
Lagrange multipliers lT,

I(q̂, a) = J(q̂, a) - lTL(q̂, a) (27)

and proceeds along the same lines to define the sensitivity of the augmented func-
tional to small parameter changes. Either way, the objective of adjoint-based
optimization is the identification of values for the variables a which minimize the
nonlinear objective function subject to Eq. (23). Local minima for the design
variables are sought, typically by a steepest-descent method which defines the
increment Da of the design parameter in terms of the sensitivity dJ/da evaluated
in Eq. (24) and a small step-size parameter e through

dJ
Da = -e (28)
da

Given an arbitrary number of design parameters, a, the local minima thus


attained represent the solution which satisfies the constrained adjoint-based opti-
mization problem. Note that the time dependence of the small-amplitude pertur-
bations may be incorporated into the adjoint-based optimization scheme either as
an eigenvalue-problem or an initial-value-problem, respectively resulting in con-
trol of modal or non-modal perturbations. Note also that the ideas discussed here
in a linearized Navier–Stokes equation context are applicable to the much wider
studied shape-optimization issue, the latter using the appropriate lower-level (e.g.,
Euler) equations; this task is accomplished by formal substitution of the linearized
operator L appearing in Eqs. (23), (25) and (27) by the nonlinear operator, N,
pertinent to the appropriate (i.e., two- or three-dimensional, incompressible, or
compressible) equations of motion (Giles and Pierce, 2001a). The interested
reader is referred to the recent review article by Kim and Bewley (2007) for further
details on the theory of adjoint-based flow control.

C. Linear State-Space Representation of Control Systems


The second approach to theoretical flow control studies is linear-systems theory
(Kim and Bewley, 2007). In particular, the state–space representation of a finite-
dimensional time-invariant (autonomous) continuous-time linear dynamical sys-
tem is expressed as a system of two ODE:

du
= Au(t ) + Bv(t ) (29)
dt

y(t) = Cu(t) + Dv(t) (30)


ROLE OF INSTABILITY THEORY IN FLOW CONTROL 87

Fig. 3 Schematic representation of the linear state-space loop (29–30) in full lines; in
addition, the feedback control loop elaborated upon in other chapters of this volume
is added by dashed lines.

The vectors u, v and y respectively represent the state, input, and output of the
state–space model, composed of the state equation (29) and the output equation
(30). The matrices A, B, C and D are respectively known as the state, input, out-
put, and direct transmission matrices and are taken to be independent of time;
usually one takes D  0. The situation is schematically depicted as the loop shown
by full lines.
The general solution of the dynamic equations (29–30) is

u(t ) = e At u 0 + Ú e A ( t -s ) Bv (s)ds (31)


0

y(t ) = Ce At u 0 + Ú Ce A ( t -s ) Bv(s) ds + Dv(t ) (32)


0

If the system that is to be controlled in question is nonlinear, linearization pre-


cedes application of the linear state–space representation (29–30). This provides the
connection with the equations of fluid motion: linearization about a background
flow is the step preceding application of flow control ideas based on the linear state–
space representation. In this case the pair (A, u) retain the physical significance of
the operator describing the linearized Navier–Stokes and continuity equations and
vector of small-amplitude perturbations, as introduced in the previous sections. The
particular case of no-control may be expressed by taking B = C = 0 in the system
(29–30) and represents the instability analysis discussed earlier. Since what follows
is equally applicable to the total flowfield, though, hats have been omitted from the
notation. Discussion of the significant issue of controllability of the system* is omit-
ted, but may be found in the recent article of Kim and Bewley (2007).

*The eigenvalues of the matrix corresponding to a controllable system may be moved at will on the

complex plane.
88 V. THEOFILIS

D. On the Numerical Solution of the Linear Control Problem


for Large Systems
At this point two classes of control problems may be introduced, on the one
hand feedback control and the Linear Quadratic Regulator (LQR) problem and on
the other hand stochastic estimation and the LQG problem. In order to avoid
duplications, the interested reader is referred to the next chapter of this book,
where elements of flow control theory are exposed. Here, some details are pre-
sented on a general method, known as linear quadratic optimization or the LQR
problem, dealing with optimization of the performance of a deterministic system:
given matrices Q and R, find a control signal v(t) such that the quadratic cost
function

J QR (u ) = Ú ÈÎu T Qu + vT Rv˘˚ dt (33)


0

is minimized, subject to

du
= Au(t ) + Bv(t ), u(t = 0) = u 0 (34)
dt

The matrices Q and R, which are used as weights for the state u and control v
vectors, respectively, may be used to construct the quadratic forms

uTQu and vTRv (35)


respectively representing the deviation of the system from its initial state u0 and
the cost of the control scheme. According to the (continuous) LQR theorem, the
constrained minimization problem of Eqs. (33–34) has a unique solution, U,
obtained by solving the algebraic Riccati equation

UA + ATU + Q - USU = 0 (36)

with S = BR-1BT. This solution may be used in order to construct the control gain
matrix

K = R-1BT U (37)

and, in turn, the vector v0(t) = -Ku(t); the LQR theorem asserts that v0 is the unique
optimal control vector which minimizes the functional JQR defined in Eq. (33).
Numerical solution of the algebraic Riccati equation

UA + AT U + Q - UBR-1BT U = 0 (38)

is a key element of the flow control problem. As such, stable numerical methods
for the solution of the Riccati equation have been the subject of intense investiga-
tion over recent decades. Assuming a modest leading dimension, n, of the matrices
involved in Eq. (38), direct solution methods exist which can be broadly classified
into invariant subspace, deflating subspace, and Newton methods (Datta, 2004).
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 89

An appropriate method of this class would come into play in flow-control prob-
lems involving the numerical discretization of a single spatial direction, whereas
the other two directions are taken as homogeneous and expanded by a Fourier
Ansatz. This is the case of most flow control applications considered so far (e.g.,
plane channel and single vortices), as will be discussed in the next section.
However, in a manner analogous to that exposed earlier regarding (Bi- and Tri-
Global) flow instability, flow control applications in complex geometries involve
coupled resolution of two (or three) spatial directions. In that case the aforemen-
tioned methodologies for direct numerical solution of the matrix equations/systems
are impractical, since such methodologies scale like O(N 3), N being the leading
dimension of the discretized matrix. On the other hand, independently of the dis-
cretization method used, coupling multiple spatial directions involves the intro-
duction of a large degree of sparsity in the numerical problem to be solved. The
inefficiency of the direct methods stems precisely from the large number of zeroes
introduced by the coupled spatial discretization. Again, by analogy to instability
theory (Theofilis, 2003), the method of choice currently in use for this class of
problems relies on Krylov subspace iteration (Arnoldi, 1951), its key advantage
over direct methods being that subspace iteration involves matrix–vector multipli-
cations, which may be performed while preserving the sparsity of the large-scale
problem. The approach for the solution of a flow control problem would be to use
subspace iteration to project the original large-scale problem onto a Krylov sub-
space and solve the projected problem in this subspace using direct techniques.
As an example, the Krylov subspace iteration method proposed for the algebraic
Riccati equation (38) by Jaimoukha and Kasenally (1994) solves
GH + HTG - GBBTG + LLT = 0 (39)

for G, an estimate of U. Here, for simplicity, R = I and Q = LLT have been taken
and, crucially, H is the Hessenberg matrix of dimension m << n generated by a
block Arnoldi algorithm (Saad, 1996). An estimate of the desired solution is then
reconstructed by
U = MGMT (40)

where M is the matrix of vectors generated as part of the same Krylov subspace
iteration process.
An alternative technique for the numerical solution of the Riccati equation has
been followed by Hoepffner et al. (2005), who employed the Kailath implementation
of the Chandrasekhar method, which reduces the O(N2) cost of solving the Riccati
equation to O(c * N) operations, N being the leading dimension of the system and the
constant being related to the number of inputs and/or observations (Kailath, 1973).
This approach has been found to work well for the parallel channel flow addressed
by Hoepffner et al. (2005), but has not been tested on larger systems.
In the face of the numerical challenges that the solution of the Riccati equation
poses, it is worth mentioning the model reduction efforts aimed at reducing the
size of the problem solved prior to applying flow control; the interested reader
is referred to Rowley (2005), Åkervik et al. (2007) and references therein, as well
as the next chapter of this monograph for an introduction to this issue and further
references for this rapidly growing field.
90 V. THEOFILIS

IV. Applications
The application of the ideas exposed in the previous section is made a little
more explicit in this chapter. By way of introduction, it is noted that three distinct
paths to flow control by (passive or active) control of flow instabilities may be
taken. First, experimentation may be used in order to understand a specific prob-
lem and, on the basis of modifications achieved on a canonical geometry, build
intuition for problems of the same class. In this context, it is unclear (and, up to a
certain point, immaterial) whether the modifications achieved are the result of
modifications of the instability characteristics of the flow. Some of the most suc-
cessful applications of flow control belong to this category; probably the most
cited example in this context is the observation of Prandtl (1904) that separation
may be delayed through suction. The work of Wygnanski and co-workers
(Greenblatt and Wygnanski, 2000; Seifert and Pack, 1999) also falls in the same
category, although these authors have conjectured that their successful control of
separation is the result of control of an instability mechanism.
The second approach is where instability information is available and exploited
for flow control purposes. Early successes have been reported by various authors,
notably in actively controlling two-dimensional Tollmien–Schlichting (TS) waves
by the introduction of anti-phase perturbations (Milling, 1981; Schilz, 1965;
Thomas, 1983). The same well-studied example of TS-wave instability also serves
to highlight the perils of extrapolation of a flow control methodology beyond its
design parameters in the absence of full knowledge of the instability physics. If
flow control via control of specific flow instabilities were to be followed, full
knowledge of the instability physics may not be sufficient but is certainly a neces-
sary condition for success. Indeed, while the compressible analogue of TS waves
is stabilized by suction and/or wall cooling, high-speed BL flow would not be
relaminarized by either of these (passive) flow control approaches. The reason is
the existence of additional eigenmodes of high-speed BL flow, discovered theo-
retically by Mack (1984) and verified experimentally in subsequent works
(Lysenko and Maslov, 1984), which respond to suction and wall cooling in exactly
the opposite manner to that of TS waves. Further, in the course of subsequent
flow-control investigations, it becomes clear that monitoring a single eigenmode
may well have the desired effect on the target perturbation but does not exclude
the possibility of (potentially catastrophic) generation of other instabilities of the
modified system, which would themselves have to be controlled. In other words,
a line of thinking currently exists according to which the availability and exploi-
tation of full eigenspectrum information does not guarantee the controllability of
the system.
Contrasting with the latter view are the spectacular successes of both passive
and active control of systems in which instability theory plays a central role, such
as control of crossflow vortices by distributed roughness elements (DRE) on
crossflow-instability dominated swept wings (Saric et al., 1998) and hypersonic
flow control using ultrasonically absorptive coatings (UAC) (Fedorov et al., 2001,
2003). Behind DREs is the idea of passive modification of the background flow
characteristics in a way in which the modified flow transitions later than the base-
line although a flow results which is linearly more unstable than the baseline
configuration. The existence of such modified states has been discovered experi-
mentally and confirmed through nonlinear PSE calculations (Haynes and Reed,
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 91

2000); the objective then becomes optimization of the placement of the DREs at
different conditions in order to maximize the extent of laminar flow on the wing
surface. Interestingly, this technology has recently been demonstrated also to be
effective in supersonic flow (Saric and Reed, 2002). Turning to the hypersonic
flow regime, where active and reactive laminar flow control (LFC) methods are
not practical under the severe conditions present, passive LFC concepts are needed
to delay transition on hypersonic vehicle surfaces. In contrast to re-entry con-
figurations, local Mach numbers Me on the hypersonic air-breather surfaces are
relatively large. This leads to the following significant changes: transition due to
roughness becomes ineffective since critical roughness height quickly increases
with Me; the first mode is stable because of low wall temperature ratios; 2-dimen-
sional shaping of hypersonic air-breathers diminishes cross-flow instability.
Under these conditions the second mode is a dominant instability leading to tran-
sition. The second mode results from an inviscid instability driven by a region of
supersonic mean flow relative to the disturbance phase velocity. This mode
belongs to the family of trapped acoustic modes propagating in a wave-guide
between the wall and the sonic line (Gushchin and Fedorov, 1989; Mack, 1984).
Fedorov et al. (2001) assumed that a passive UAC of fine porosity may suppress
the second mode by a disturbance energy extraction mechanism and at the same
time may not trip the boundary layer due to roughness effects. This UAC–LFC
concept was confirmed by the linear stability analysis (Mack, 1984) and by mea-
surements of transition loci (Gushchin and Fedorov et al., 2001) and BL distur-
bances (Fedorov et al., 2006; Rasheed et al., 2002) in hypersonic wind tunnels on
sharp cones with UAC of regular structure (equally spaced cylindrical blind
microholes) and random structure (felt-metal coating). The nonlinear aspects of
UAC with regular microstructure were studied in the experiment of Fedorov et al.
(2003). It was shown that the harmonic resonance, quite pronounced in the distur-
bance evolution on the solid surface, is completely suppressed on the porous
surface. Futher, the degree of nonlinear phase locking which is associated with
the sub-harmonic resonance and identified on the solid surface is substantially
weakened on the porous surface. These studies confirm the immense technologi-
cal importance of UAC. Note that a thin passive porous coating can be naturally
integrated into thermal protection systems with virtually no penalties.
Theoretical knowledge of instabilities may be used in order to devise passive
flow control methodologies based on exploitation of the transient growth pheno-
menon and in this context it is worth mentioning the ideas of Cossu and Brandt
(2002) and Fransson et al. (2005). In a manner analogous to the ideas underlying
the DREs, these authors exploit linear instabilities in order to modify (stabilize) the
basic BL flow and excitation is also achieved by appropriate placement of rough-
ness elements in the leading-edge region. The essential difference with DREs lies
in the mechanism used for excitation; unlike the modal perturbations used in the
DREs, non-modal optimal perturbations (streamwise streaks) are excited in the
context of the works of Cossu and Brandt (2002) and Fransson et al. (2005).
In the aforementioned classes of studies it is often found that advance knowledge
of the underlying flow instability physics may lead to successful (passive) flow
control methodologies. Further discussion of these will not be presented here, first
since they are discussed elsewhere in this volume and, second, in order to make
room to elaborate upon a third class of methodologies, namely adjoint-based flow
92 V. THEOFILIS

control, including both receptivity and linear-systems approaches. Minimal refer-


ence will also be made to the wide body of flow control work via reduced-order-
modeling, a technique which is essential to deal with very large systems such as
those resulting from the spatial discretization of a multidimensional problem; the
interested reader is referred to the next chapter for details on reduced-order model-
ing. Here it suffices to mention that theoretical evidence exists, according to which
the incorporation of linear instability results (eigenmodes) significantly augments
the performance of otherwise successful reduced-order-model approaches; see
Noack et al. (2004) and subsequent work by this group for details.
Turning to adjoint-based optimization, this approach has been introduced into
fluid mechanics research in the context of full-configuration optimization
(Jameson, 1989) and has since been applied to control flat-plate and channel BL
transition (Joshi et al., 1997; Joslin et al., 1995b), optimize suction on aircraft
wing sections (Pralits and Hanifi, 2003), modify the wake of non-parallel flows
such as the circular cylinder (Hill, 1992), cavity-induced laminar separation
(Hoepffner et al., 2007), and control noise amplitude in compressible turbulent
jet flow (Wei and Freund, 2006), to name but a few examples. Adjoint-based
methods are presently most commonly utilized to active flow control as they have
the potential to control the flow without prior knowledge of the detailed instabil-
ity physics although (in the context presented here) adjoint-based optimization
does act upon linear flow disturbances at their birth. The exact result obtained
depends on the choice of the objective function optimized; using the analogy
with external aerodynamics, rather than optimizing integral quantities such as lift
and drag, in the problem at hand, it is typically the energy of the linear perturba-
tions that need to be minimized. If nonlinear phenomena are not predominant,
the equations to be solved alongside the optimization of the objective function
are the direct and adjoint LNSE; otherwise, a DNS-based methodology must be
used, although the cost of the latter approach may be prohibitively high for para-
metric studies (Joslin, 2001). Note that in the case of adjoint-based optimization
using DNS, appropriate definition of the objective function permits using the
same algorithm for control of both transitional and turbulent flows (Bewley and
Liu, 1998; Bewley et al., 2001; Chevalier et al., 2002; Collis et al., 2000). In this
section, the classification of Table 1 is followed in order to present the direct and
adjoint linear instability operators introduced earlier, discuss key properties of
the equation systems to be solved, and briefly comment on the results obtained in
selected representative applications.

A. 0-D: No Flow—Natural Convection


Starting with the simplest case of no flow, i.e., (ū1, ū2, ū3)T  0, which is never-
theless interesting in the analysis and control of phenomena such as Rayleigh–
Taylor instabilities of plasmas, one notes that the associated linearized operator is
normal.* The consequence of this property is that, in this simplest of situations,
the only relevant limit for instability is that of the long-time behavior of the

*Flow at constant speed is equivalent, by a Galilean transformation, to this problem.


ROLE OF INSTABILITY THEORY IN FLOW CONTROL 93

solutions (2) to the initial-value-problem (1). Solution of eigenvalue problem


resulting by the substitution

du
∫ iw u (41)
dt

for the determination of a complex eigenvalue w suffices for the prediction of the
instability properties of the system in question and, in turn, of the eigenmodes
which will be focused on in a flow control context. Instability problems of this
class are the subject of active research using the modal concept (Thiele et al.,
2006). Caution is advised, however, that the normality of the operator describing
linear instability does not necessarily translate in to the normality of the operator
describing linear control (Zhao and Bau, 2006), such that the generalized stability
theory presented earlier becomes relevant.

B. 1-D: Parallel Background Flows


In the context of one-dimensional basic or mean flows dependent upon one
spatial coordinate, say x2, the parallel-flow assumption must be made, implying
that the basic flow velocity component ū2 along the spatial coordinate x2, vanishes.
Note that x2 can be the spatial direction along the wall-normal, x2  y, e.g., in a
free-shear- or boundary-layer flow context or along the radial spatial direction,
x2  r, when referring to an axisymmetric problem, e.g., instability of the flow-
field around an isolated vortex, or of Taylor–Couette flow between concentric
cylinders. Owing to the analytical simplicity and numerical tractability of the
resulting equations describing both linear stability and linear control, this class of
problems has been used as a testbed of many modern ideas exposed in the previ-
ous sections.
In the case of a plane shear flow comprising a single-component basic flow
(a “profile”), u = (ū1( x2), 0, 0)T = (ū1( y), 0, 0)T linearized equations of motion
may be written for the determination of the amplitude functions of the distur-
bance velocity components and the disturbance pressure. In the absence of
viscosity the disturbance pressure may be eliminated from the linearized equa-
tions of motion leading to the (one-dimensional) Rayleigh equation which
governs inviscid linear flow instability; see classical monographs for details
(Drazin and Reid, 1981). The strong (inviscid) instability of the underlying
background flow may be exploited in order to control both laminar and turbu-
lent flow; the most-cited example of this class is the turbulent mixing-layer,
where control is achieved by excitation of its Kelvin–Helmholtz instability
(Greenblatt and Wygnanski, 2000; Oster and Wygnanski, 1982). Of further
interest presently is the fact that the pertinent linearized operator is non-normal.
As such, besides the inviscid modal instability predicted by solving the
Rayleigh equation, potential for transient growth exists. Precise knowledge of
both modal and non-modal perturbations would be necessary if active flow
control were attempted on the basis of exploitation of instability characteris-
tics. By contrast, as has been mentioned elsewhere, adjoint-based optimal
control does not require this information, although it does act to compensate
the action of the underlying linear operators.
94 V. THEOFILIS

Reinstating viscosity in the problem, the LNSE in a Cartesian frame of refer-


ence read

du (42)
= Au
dt

for the determination of u = (û2, ẑ 2)T, the components of the vector denoting
disturbance velocity and disturbance vorticity components, along the y spatial
coordinate. Here A = M-1L with

Ê LOS 0 ˆ
Êa 2 + b 2 - d 2/dy 2 0ˆ
L = Á du1 ˜, M=Á (43)
Á ib LSQ ˜ Ë 0 I ˜¯
Ë dy ¯

with the parameters a and b defined in Table 1. Initial conditions on û2 and


zˆ2 alongside the boundary conditions û2 = dû2/dy = zˆ2 = 0 at solid walls and/or
the free-stream complete this system. The operator M being positive definite,
the behavior of the initial-value problem (42) is determined by the properties
of the operator L which, in turn, is determined by the non-normality of the
Orr–Sommerfeld, LOS, and Squire, LSQ, operators,
2
1 È d2 ˘ È d2 ˘ d 2u
LOS = Í 2
- (a 2 + b 2 )˙ - ia u1 Í 2 - (a 2 + b 2 )˙ + ia 21 (44)
Re Î dy ˚ Î dy ˚ dy
1 È d2 ˘
LSQ = - Í 2
- (a 2 + b 2 )˙ + ia u1 (45)
d
Re Î y ˚

The non-orthogonality of the linearized operator suggests the need to define


the adjoint initial-value problem (Henningson and Schmid, 1992)

du*
= A* u* (46)
dt

where A* = M-1L*, u* is the vector containing the adjoint disturbance velocity and
vorticity components along y, and

Ê * d(u1 )ˆ
LOS - ib
L =Á dy ˜
*
(47)
Á ˜
Ë 0 LSQ ¯
*

LOS
*
, and LSQ
*
being the adjoint Orr–Sommerfeld and Squire operators. These
definitions, in conjunction with Eq. (6) permit identification of both modal (in the
limit t Æ •) and non-modal (at finite-times) instabilities. In turn, methodologies
can be devised in order to control the disturbances prevailing at the appropriate
parameter limits.
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 95

At this point it is appropriate to return to the two predominant paths to


theoretically-founded adjoint-based flow control efforts of one-dimensional back-
ground flows, introduced in the previous two sections. The first is associated with
the solution of the receptivity problem and consequent identification of regions of
high sensitivity in the flow, to be exploited in a flow-control context. Although the
first application of adjoint equations to study the receptivity problem in a bound-
ary layer is often attributed to Hill (1995), it was the work of Tumin and Fedorov
(1984) which first successfully employed this concept in the context of linear
instability and receptivity analysis of the weakly-non-parallel flat-plate BL; the
same approach has subsequently been used to study receptivity of the strictly
parallel Hagen–Poiseuille flow (Tumin, 1996). The second analysis path is a linear-
system approach which employs the LNSE and/or full (two- or three-dimensional)
direct numerical simulation, without regard to the particular class of flow instabil-
ity that is being controlled. The groundbreaking studies of Joslin and co-workers
(Joslin, 2001; Joslin et al., 1995b, 1997) were the first of this class, followed by
the works of Joshi et al. (1997) and Bewley and Liu (1998) on channel flows and
the work of Gallaire et al. (2004) on control of instabilities of an isolated vortex.
Before discussing these works, it is interesting to stress that the problems (42)
and (46), as well as their linear-control counterparts based on Eqs. (29–30), or the
systems resulting from an LQR or an LQG formulation of the control problem,
are all ODE based. In this respect, numerical techniques for the solution of both
the linear flow instability and linear flow control problems may be based on classic
dense-matrix algebra techniques (Golub and van Loan, 1996). However, even for
the one-dimensional problems involved, further reduction of the instability prob-
lem from a system to a single-equation is interesting from a numerical effciency
point of view; one such reduction is discussed by Joshi et al. (1997).
The numerical tractability of the optimization and control problems associated
with one-dimensional background flows has resulted in them having been used
over the years as testbeds for the application of the ideas exposed herein. Tumin
and Fedorov (1984) were the first to recognize the need to complete the theory of
amplification of linear perturbations in a flat-plate boundary layer with an asso-
ciated theory of receptivity.* They went on to demonstrate how the use of bi-
orthogonal sets of direct and adjoint eigenfunctions provides means of calculation
of the amplitudes of the resulting instabilities and information on the spatial loca-
tions where these amplitudes are maximized; one such result is shown in Fig. 4.
An adjoint technique has also been used in the influential work of Hill (1995),
who studied TS-excitation via a wide variety of sources, associating in all cases
the amplitude of the adjoint TS-perturbations with the value of the receptivity
coefficient. In a series of papers, Joslin and co-workers (Joslin, 2001; Joslin et al.,
1995b, 1997) used two-dimensional direct numerical simulation to revisit the
problem of active flow control of BL instabilities, schematically presented in
Fig. 5. Two distinct flow control paths were followed by these authors, first
employment of the by then well established wave-cancellation (WC) technique
(Joslin et al., 1995b), and second, the introduction, for the first time in a fluid flow
instability problem concept, of linear-system control (Joslin et al., 1997). Figure 6

*A theory which these authors called “susceptibility to external effects.”


96 V. THEOFILIS

Fig. 4 Adjoint based predictions (curve 1) against measurements (curve 2) of


TS-wave amplitude, introduced by low-frequency ( f = 70 Hz) oscillations of a vibrat-
ing ribbon at x = 510 mm (Tumin and Fedorov, 1984).

shows results of the first approach, which demonstrates that WC may effectively
damp TS waves, linear amplification in the controlled case starting downstream of
the location where growth would be expected according to linear theory. In the
same figure results of the adjoint-based control are displayed; it may be appreci-
ated that the (targeted) WC approach performs better than adjoint-based optimiza-
tion, although the latter approach has the obvious benefit of being oblivious to the
instability it controls.
While the adjoint-based optimization work of Joslin et al. (1997) employed
DNS, both Joshi et al. (1997) and Bewley and Liu (1998) used linear-systems
theory based on the LNSE in incompressible flow and relatively simple geom-
etries; the work of Collis and Lele (1999) has pioneered the use of the LNSE in
compressible flow around a swept cylinder. Joshi and co-workers. have stressed
that this approach shifts the flow control problem away from thinking about
how inputs can mitigate a linearly unstable system, e.g., by employing the WC
concept, to thinking about how sensors and actuators can be incorporated in the
feedback loop of the LQR class in order to stabilize a system. Figure 7 shows
the effect of control in two-dimensional 4p-periodic plane channel flow at the

Fig. 5 AFC concept for stabilization of linear disturbances (Joslin et al., 1995).
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 97
LIVE GRAPH LIVE GRAPH
Click here to view
Click here to view

Fig. 6 a) Wave-cancellation applied to Blasius boundary layer via two distinct methods
acting on the wall-normal velocity component vw and compared with the no-control
case vw = 0 (Joslin et al., 1995b). b) Comparison of uncontrolled, WC-controlled, and
adjoint-based controlled unstable flat-plate flow (Joslin et al., 1997).

linearly-unstable conditions Re = 10000; a = 1. Applying control after 200


time units virtually eliminates the unsteadiness associated with the modal
growth. Interestingly, at subcritical conditions, where one path to transition
would be initiated by linear amplification of three-dimensional perturbations,
control not only damps the energy of the imposed disturbance (effectively
stabilizing the flow faster), but also damps the growth of the three-dimensional
disturbance, thus further delaying transition. Analogous experience and results
were obtained in the work of Bewley and Liu (1998); the key difference of this
to all prior works is that it dealt with optimal (also known as LQG or H2) and
LIVE GRAPH
Click here to view LIVE GRAPH
Click here to view

Fig. 7 a) Various measures of the effectiveness of control in plane channel flow (Joshi
et al., 1997) at unstable conditions. b) Development of energy of the imposed two-
dimensional and the generated three-dimensional perturbations without (solid) and
with (dashed) control at subcritical conditions.
98 V. THEOFILIS

Fig. 8 Reduction of maximum transient energy in subcritical channel flow (Bewley


and Liu, 1998). Upper to lower curve: no-control, by application of proportional
control (Joshi et al., 1997) and two optimal control methodologies.

robust (H•) control of the non-normal linear system at hand, introducing the
concept of appropriate transfer functions to characterize the feedback-control
approach used. In so doing, the approach proposed is capable of dealing with
both modal and non-modal instabilities, both being relevant to shear-flow
transition (Schmid and Henningson, 2001). Focusing on the channel flow
example at subcritical conditions and using the LNSE, Fig. 8 shows the damp-
ing of transient growth energy with time, when the proportional control
approach of Joshi et al. (1997) and two different optimal control methodolo-
gies proposed by Bewley and Liu (1998) are applied. On the other hand, Collis
and Lele (1999) have chosen to work on a swept cylinder model, which permits
studying curvature and non-parallelism effects both on receptivity and instabil-
ity. It was found that these effects counteract each other in terms of receptivity,
while the strong modication of parallel-flow predictions underlined the neces-
sity to relax the restrictive assumprions of the latter theory when dealing with
realistic flow geometries.
Turning to vortical flows, to date there exists a single theoretical flow control
work by Gallaire et al. (2004), who have recently applied the tools exposed in
this chapter to control an isolated confined vortex. This problem belongs to a
wide class of applications involving vortical flows, which have mainly received
experimental attention. The substantially more complex problem (owing to the
open and developing nature of the background flow) vortex breakdown on a
delta wing has been the subject of both open- (Vorobieff and Rockwell, 1996)
and closed-loop (Gursul et al., 1995) experimental flow control studies.
Interestingly, and in a manner analogous to periodic excitation of separated
flow (discussed in detail elsewhere in the present volume) Vorobieff and
Rockwell (1996) have shown that unsteady blowing at the trailing edge of a
delta wing is an effective open-loop mechanism delaying vortex breakdown,
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 99

LIVE GRAPH
Click here to view

Fig. 9 Dependence of perturbation energy on time in a linearly unstable confined


vortex flow at a given set of swirl and control parameters (Gallaire et al., 2004). Solid
line corresponds to the baseline case, dashed line to the closed-loop controlled flow,
and dotted line indicates the energy necessary to stabilize the flow.

whereas Gursul et al. (1995) have designed experimentally a closed-loop con-


trol scheme capable of ensuring maintenance of the position of vortex break-
down under changes in angle of attack. On the other hand, one of the novelties
in the theoretical work of Gallaire et al. (2004) has been the implementation of
a reduced-order-model based on a small number of the least stable flow eigen-
modes and a closed-loop flow control approach to study active feedback flow
control of the confined vortex problem. The solution of the Riccati equation
(36) was used in order to construct the gain matrix associated with the feedback-
control problem. Results obtained showed mitigation of linear eigenmode
growth and, under certain (realistic) conditions stabilization of the most dan-
gerous flow eigenmode; Fig. 9 shows the latter result.

C. 1.5-D: Developing Background Flows


A major breakthrough in linear and nonlinear instability analysis has been the
introduction of the concept of parabolized stability equations (PSE); see Herbert
(1997) for a review. Unlike the initial-value- or eigenvalue-problem techniques
used for parallel flows, PSE may address weakly non-parallel flows by incorpo-
rating through a consistent inverse-Reynolds-number expansion into the insta-
bility analysis the wall-normal basic flow velocity component, which is neglected
in local linear analysis; further, PSE can treat both linear and nonlinear flow
instability. Analysis demonstrates the near-parabolic nature of the disturbance
equations (Li and Malik, 1997) which may be solved by space-marching tech-
niques; the same numerical approach may be used in order to solve the adjoint
PSE equations.
100 V. THEOFILIS

The direct PSE equations read

Ê ˆ
L + u- x + v- d/dy u- y 0 ia Ê uˆ ˆ
Á 1D ˜
Á - - - ˜ Á vˆ ˜
Á vx L1D + v y + v d/dy 0 d/dy˜ Á ˜
Á - - ˜ Á wˆ ˜
Á w x w y L1D + v- d/dy ib ˜ ÁË pˆ ¯˜
ÁË ia d/dy ib 0 ˜¯
Ê - 2ia ˆ
u- 0 0 1 (48)
Á Re ˜
Á ˜ Ê uˆ x ˆ
2ia Á
Á 0 u- - 0 0˜ vˆx ˜
+Á Re ˜ Á ˜ =0
Á ˜ Á wˆ x ˜
2ia
Á 0 0 u- - 0˜ Á
Ë pˆ x ˜¯
Á Re ˜
ÁË 1 0 0 0˜¯

where L1D = -iw + i + i(aū + bw̄) + 1/Re(a2 + b 2 - d 2/dy2) is an one-dimensional


linear operator, in the resolved spatial direction, y  x2. If a harmonic dependence
of the perturbation amplitude functions q̂ = (û, v̂, ŵ, p̂)T with the downstream spa-
tial direction, x1, is assumed, the operators L1D in (48) and LOS and LSQ in (43) can
be related to one another.
The key characteristic of the PSE is the existence of two length scales, one
slow upon which the background flow develops and one fast upon which the
amplitude functions of the instability waves depend; a normalization condition
is used in order to uniquely define this scale separation (Herbert, 1997). The
results of linear PSE, as described by Eq. (48), are in close agreement with
those of local instability analysis (strictly applicable to parallel flows alone),
while essential factors such as the non-parallel nature of the basic state (dem-
onstrated by the presence of a basic flow velocity component, ū2 = v̄, along the
wall-normal direction, y, in Eq. (48)) permits a more accurate description of
instability physics by PSE, compared to that provided by the local linear
(Orr–Sommerfeld- and Squire-equation-based) approach. Probably more sig-
nificant, though, is the fact that the results of nonlinear PSE are in excellent
agreement with those of spatial DNS (Bertolotti et al., 1992), whereas the
computational cost of nonlinear PSE is orders of magnitude lower than that of
spatial DNS. The latter property is a consequence of the fact that the system in
Eq. (48) may be solved by spatial marching along the streamwise direction,
x  x1, while at each x1 location an one-dimensional problem along the wall-
normal y  x2 spatial direction may be solved; the starting conditions for the
marching are provided by solution of an Orr–Sommerfeld problem. It is pre-
cisely the efficient nature of the PSE solution procedure that makes this analysis
methodology an ideal candidate on which flow control via adjoint-based-
optimization may be developed.
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 101

The adjoint system to which the direct PSE problem (48) is coupled reads
(Dobrinsky, 2002)

Ê * - v- ˆ
L - u x + d/dy - v- x -w -
x ia * Ê uˆ ˆ
Á 1D ˜
Á - v
- v - - ˜ Á vˆ ˜
Á - u y L*
1D - y + d/dy - w y 0 ˜ Á ˜
Á ˜ Á wˆ ˜
Á 0 0 L1*D + v d/dy -ib ˜
-
ÁË pˆ ˜¯
ÁË ia * d/dy -ib 0 ˜¯
Ê - 2ia * ˆ
Á u - Re 0 0 1˜ (49)
Á ˜ Ê uˆ x ˆ
Á 2ia *
0 u- - 0 0˜˜ Á vˆx ˜
+Á Re Á ˜ =0
Á ˜ Á wˆ x ˜
Á 2ia *
0 0 u- - 0˜ Á
Ë pˆ x ˜¯
Á Re ˜
Á ˜
Ë 1 0 0 0¯

where L1D *
= iw + i(a *ū - bw̄) + 1/Re[(a *)2 + b 2 - d 2/dy2] and, typically, a * = -a.
Appropriate normalization conditions and objective functions related to the pertur-
bation energy close the instability and flow-control problems, respectively. The
interested reader is referred to the original works for further technical details.
The first mention of the adjoint PSE concept in a flow control context has been by
Herbert (1997); without being explicit on the equations solved, this work discussed
application of the adjoint PSE to study receptivity of incompressible flow on a swept
cylinder (Poll, 1985). Collis and Dobrinsky (2000), Dobrinsky (2002) and Dobrinsky
and Collis (2000) introduced a general framework for both the direct and the adjoint
compressible Navier–Stokes equations, from which the direct and adjoint PSE sys-
tems in Eqs. (48–49) were derived. Comparisons of predictions of the adjoint PSE
and the full adjoint Navier–Stokes equations delivered an equally good agreement
as that of their direct counterparts (Bertolotti et al., 1992). These authors then went
on to perform adjoint-based receptivity studies of the Blasius and the Falkner–Skan
flat-plate boundary layers and demonstrated the very close agreement of the results
obtained and those of the previous adjoint local analysis of Hill (1995). Figure 10
shows adjoint-based receptivity results for the Blasius BL at different values of the
non-dimensional frequency parameter F. Interestingly, adjoint-based receptivity
analyses of pressure gradient effects on receptivity showed a competition between
instability and receptivity; when amplification rates were strong (e.g., by increasing
the adverse pressure gradient) the receptivity coefficients decayed and vice versa. In
parallel developments, Airiau et al. (2000), Pralits et al. (2000), and Walther et al.
(2001) have also presented receptivity prediction methodologies based on direct and
adjoint PSE equations. The latter work focused on receptivity of the (incompress-
ible, non-parallel) Blasius BL and documented the response of this flow to wall
forcing via roughness, vibration, and transpiration, as well as acoustic receptivity
LIVE GRAPH
Click here to view
102 V. THEOFILIS

Fig. 10 Normalized amplitudes of the adjoint perturbations resulting from differ-


ent means of forcing the Blasius boundary layer: a) and b) maximum receptivity to
point forcing and corresponding distance from the wall, c) receptivity to suction/
blowing at the wall, and d) receptivity to streamwise velocity disturbances (Collis
and Dobrinsky, 2000).

via scattering of incident acoustic waves on a hump, following the approach intro-
duced by Hill (1995); of particular interst has been the very good agreement of the
direct/djoint PSE calculations of Airiau and co-worker. with the local theory results
of Crouch (1992). The direct/adjoint PSE methodology proposed was shown to be
effective in increasing the Reynolds number at which a TS wave of a particular
frequency would become unstable; these results are shown in Fig. 11.
At this point it is appropriate to introduce the related work of Luchini and Bottaro
(1998), who employed a direct/adjoint methodology based on the parabolic BL
equations in order to study receptivity of Görtler vortices. As has repeatedly been
mentioned, the motivation for using an adjoint formulation in this context is the
avoidance of the need to cover a wide parameter space by a large amount of direct
analyses. Instead, these authors solved the receptivity problem by appropriate
Green’s functions, calculated via parabolic adjoint equations of the same class as
Eq. (49). A key contribution of the work of Luchini and Bottaro (1998) has been
the reintroduction into the instability, receptivity, and control literature of the con-
cept of “optimal control,” first presented in the work of Hill (1992) on the control
of the (non-parallel) wake of the (non-parallel) circular cylinder flow. This influen-
tial work has given rise to several subsequent studies which dealt with the problem
of instability, receptivity, and control of disturbances in BL flow employing analo-
gous techniques; see the recent work by Zuccher et al. (2006) and references
therein. Here two of the earliest investigations which employed the direct/adjoint
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 103

LIVE GRAPH
Click here to view

Fig. 11 Increase of the Reynolds number at which a TS-wave of a particular fre-


quency, here F = 50, becomes unstable; shown is the baseline uncontrolled flow (solid
line) and two different control cases (dashed lines) (Walther et al., 2001).

optimization method proposed by Luchini and Bottaro (1998) are briefly discussed,
namely the work of Corbett and Bottaro (2001) on optimal perturbations of
Falkner–Skan–Cooke flows, and that of Cathalifaud and Luchini (2000) on control
of optimal perturbations in flat and concave-wall BL flows. Corbett and Bottaro
(2001) used direct/adjoint operators closely related to those presented in Eqs. (43)
and (47) and a disturbance kinetic energy measure to close the system of optimal
perturbations. These authors document that the perturbations experiencing maxi-
mum growth are vortical perturbations evolving into streaks and attempt to establish
a connection between these and crossflow disturbances. Cathalifaud and Luchini
(2000) also performed an adjoint-based optimization analysis and concluded that
the optimal perturbations in a BL, namely those which maximize disturbance
energy growth downstream, are streamwise vortices which evolve into streamwise
streaks. Wall transpiration was used to control optimal perturbations and two
energy norms were employed, resulting in two different adjoint operators and
objective functions. Results obtained verified controllability of optimal perturba-
tions and demonstrated that there exists a direct relationship between the amount
of maximum control energy utilized an the resulting modication of the ensuing
streak perturbation; a result is shown in Fig. 12.
Turning to linear-systems approaches based on the direct and adjoint PSE,
Pralits et al. (2002), and Pralits and Hanifi (2003) have developed adjoint-based
optimization techniques for compressible flat-plate zero- and adverse-pressure-
gradient boundary layers and wing sections. The solution procedure followed is
conceptually the same as that presented in Fig. 1, although the direct/adjoint BL
equations are solved to obtain the basic states analysed, while perturbations are
obtained by solution of the direct/adjoint PSE. The objective function chosen in
the optimal control algorithm of Pralits et al. (2002), as applied to the flat-plate
BL problem, was based on controlling isolated TS waves; it was shown that this
104 V. THEOFILIS

LIVE GRAPH
Click here to view

Fig. 12 Dependence of the energy of Kelbanoff modes as a function of the control


energy, Ec (Cathalifaud and Luchini, 2000).

approach is sufficient to damp other disturbances of the same class. By contrast,


Pralits and Hanifi (2003) addressed the flow control problem on a wing section,
where TS and crossflow (CF) modes may coexist; in order to be able to capture
and control both, the objective function in this work was redefined as an integral
of disturbance energy over both the entire wall-normal and a portion of the down-
stream spatial direction. From a technological point of view an interesting result
of the latter work has been the application of the optimal control methodologies to
a wing section in which suction was realized by means of a discrete sequence of
pressure chambers. The results obtained were consistent with those of (continuous)
optimal suction and tended to the latter in the limit of large number of small-scale
pressure chambers, thus highlighting the applicability of the optimal control theo-
retical approach to practical situations.
A linear-systems approach was also applied to the (developing) Blasius and
Falkner–Skan boundary layers by three—available in the existing bibliography—
works, namely Chevalier et al. (2002, 2007) and Högberg and Henningson
(2002); in contrast to the work of Pralits et al. (2002) and Pralits and Hanifi
(2003), these authors employed both of the direct/adjoint LNSE and the full
direct/adjoint incompressible DNS equations. The feedback operator was con-
structed from the appropriate LNSE (Orr–Sommerfeld and Squire) equations,
derived under a parallel-flow assumption. This assumption permitted straight-
forward numerical solution of the associated Riccati equations (36) for each wave
LIVE GRAPH LIVE GRAPH
Click here to view Click here to view
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 105

Fig. 13 Spatially developing incompressible boundary layer: a) control of modal and


b) optimal perturbations (Högberg and Henningson, 2002).

number pair in the streamwise and spanwise directions; the same feedback was
then applied to the DNS-based feedback control of spatially developing BL flows.
Figure 13 shows the effect of DNS-based optimal control on both modal and
non-modal perturbations in a spatially developing BL, as obtained by Högberg
and Henningson (2002). The uncontrolled flow in which modal perturbations
develop is denoted by the dashed, whereas the result of flow control at three sets
of control parameters is shown by the other three line types. Stabilization of a
modally unstable flow is obtained for two of the three control-parameter sets. On
the other hand, optimal perturbations at subcritical conditions could not be elimi-
nated, although flow control resulted in the reduction of their maximum energy
growth by factor 2.* In addition, growth of steady CF vortices in a Falkner–Skan–
Cooke flow is successfully inhibited by the same controller.
In a follow-up work, Chevalier et al. (2007) applied stochastic estimation ideas to
the same problem of linear control of Falkner–Skan–Cooke flows. Rather than uti-
lizing full knowledge of the three-dimensional flowfield, as is the case when DNS
data is available (Högberg and Henningson, 1998), this work employed a stochastic
estimation system and a small set of wall-shear measurements to estimate the state.
Kalman filter theory was utilized to reduce the estimation problem to the solution of
algebraic Riccati equations of the class shown in Eq. (36). Subsequently, the estima-
tion information was fed into a spatial DNS approach which was able to control
modal (TS) and non-modal (optimal) perturbations of the spatially developing BL.
Figure 14 again shows characteristic results of controlling the two classes of insta-
bilities; despite the availability of limited and noisy information on the system,
qualitatively analogous results with those shown in Fig. 13 were obtained.
With the exception of the work of Pralits and co-workers, most of the adjoint-
based optimization and linear control work has been performed in the (largely
unexplored) incompressible flow limit. However, first applications of control of
compressible flow (instabilities) have recently appeared in the literature; besides
being a flow regime of interest in its own right, control of compressible flow
instabilities permits theoretically founded aeroacoustics research. Representative

*Here solid line denotes uncontrolled flow.


106 V. THEOFILIS
LIVE GRAPH LIVE GRAPH
Click here to view Click here to view

Fig. 14 Spatially developing incompressible boundary layer (Chevalier et al., 2007),


when limited field information is available and stochastic estimation is used to obtain
the missing data (Hoepffner, 2006): a) control of modal and b) optimal perturbations;
solid lines indicate uncontrolled flow.

of this class of efforts are the works of Collis et al. (2002), who first applied
optimal control theory to problems governed by the unsteady, compressible
Navier–Stokes equations, that of Barone and Lele (2005), who studied receptivity
of a compressible mixing layer, and a series of papers by Freund and co-workers
(e.g., Wei and Freund, 2006 and references therein) on jet-noise control. Collis
et al. (2002) presented a general formulation for the nonlinear compressible
direct and adjoint Navier–Stokes equations, alongside two different objective
functions respectively minimizing kinetic energy and wall heat transfer. Because
of the complexity of the equations to be solved, automatic differentiation tech-
niques (as opposed to explicit derivation and evaluation) were successfully
implemented; despite the efficiency of the parallel computations performed, it
was found that the part of the overall algorithm dealing with the adjoint equa-
tions was four times as expensive as that related to the direct computations. From
a realizability point of view, the concept of regularization of the spatial and tem-
poral derivatives was introduced; satisfaction with this criterion resulted in order-
of-magnitude improvements of the control variables. On the other hand, the
works of Barone and Lele (2005) and Wei and Freund (2006) dealt with control
of instabilities of compressible flows, from a receptivity and an aeroacoustics
point of view respectively. Barone and Lele (2005) posed the receptivity problem
in terms of the direct and adjoint compressible LNSE, which they solved using a
high-order compact-finite-difference-based overset method, appropriate for
aeroacoustic calculations in complex geometries. Receptivity to mass-, momen-
tum-, and heat-sources of a Mach 1.2 mixing-layer flow emanating from a finite-
width splitter plate was then computed. Interestingly, the connection of BL
modes upstream of the splitter plate with Kelvin–Helmholtz instabilities in the
mixing layer was analyzed via the adjoint equations and a mechanism for energy
transfer between the two classes of instabilities was discovered. Wei and Freund
(2006) have employed the direct and adjoint compressible LNSE in order to
perform systematic investigations of mechanisms to control radiated sound from
a fully-resolved compressible turbulent jet flow. The actuator is implemented as
a source term in a compressible DNS approach, with the objective function
being defined with the aid of pressure perturbations along a desired line in the
flowfield. It was stressed by these authors that the noise control mechanism,
which under certain conditions resulted in an 11dB reduction, was not an acoustic
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 107

cancellation* but modifications in the flow, introduced by the actuation scheme


chosen and resulting in a more uniform downstream advection of turbulent
structures.

D. 2-D: Essentially Non-Parallel Background Flows


While many applications of interest fit into the previously discussed classes,
probably most configurations appearing in practice do not afford the respective
simplifying assumptions of one- or quasi-two-dimensional background states.
Relaxing these homogeneity assumptions and treating two out of three spatial
directions as inhomogeneous permits addressing flow instability and control prob-
lems in a wide variety of configurations of engineering interest. A non-exhaustive
list focused on topics of interest to this volume includes flows in the vicinity of
geometric inhomogeneities such as wing-body junctions, rectangular and rounded
backsteps, open cavities, non-axisymmetric nozzles, axisymmetric wedge flows of
arbitrary wedge angle, shock-induced separation, and non-axisymmetric vorticity
distributions in the wake of an aircraft, besides being able to analyze all flows
treatable by the approaches of the previous two subsections without the need to
employ the assumptions made there. The price to be paid for this freedom is the
coupled discretization of the inhomogeneous spatial directions. The linearized
operator A describing instability in a BiGlobal context
du
M = Au (50)
dt
( )
I 0
with M = 0 0 and I = diag{1,1,1} can be derived in a Cartesian frame of reference
from (11) using the independence of the background state on the lateral spatial coor-
dinate z. Modal perturbations are introduced into the autonomous system according to

u = (qˆ ( x, y), pˆ ( x, y)) exp ÈÎ+i (b z - w t )˘˚ (51)

where q̂ = (û, v̂, ŵ)T and p̂ are, respectively, the vector of amplitude functions of
linear velocity and pressure perturbations, superimposed upon the two-dimen-
sional, two- (w̄  0) or three-component, q̄ = (ū, v̄, w̄)T , steady or time-averaged
background state. The spanwise wavenumber b is associated with the spanwise
periodicity length Lz through Lz = 2p/Lz. Substitution of Eq. (51) into the linear-
ized equations of motion denes the linearized operator A describing (complex)
direct LNSE in a BiGlobal context (Theofilis, 2003)
Ê ˆ
L - u- x - u- y 0 -∂ x
Á 2D ˜
Á - v- L2D - v- y 0 -∂ y ˜
x
A= Á ˜
Á - - ˜ (52)
Á - wx -w y L2D -i b ˜
Á ˜
Ë ∂x ∂y ib 0 ¯

*An analogue to the anti-phase introduction of perturbations in a TS-dominated hydrodynamic

stability problem.
108 V. THEOFILIS

where

1 Ê ∂2 ∂2 ˆ ∂ - ∂
- b 2 ˜ - u- - v - ib w
- (53)
L2 D = Á 2
+ 2

Re Ë x ∂ y ¯ ∂ x ∂ y

In an analogous manner, the quantities q̃* = (ũ *, ṽ*, w̃*)T and p̃* denote adjoint
disturbance velocity components and adjoint disturbance pressure, and the
decomposition

u* = (q ( x, y), p ( x, y))exp ÈÎ-i (b z - w t )˘˚ (54)

is introduced into the adjoint Navier–Stokes equations of motion

du*
-M = A* u* (55)
dt

Note the opposite signs of the spatial direction z and time in Eqs. (51) and (54),
denoting propagation of adjoint perturbations in the opposite direction compared
with that of the direct small-amplitude disturbances. The matrix describing the
(complex) adjoint LNSE reads

Ê † - - - ˆ
L -ux -vx - wx -∂ x
Á 2D ˜
Á - ˜
- u- y L†2 D - v- y - wy -∂ y ˜
A* = Á (56)
Á ˜
Á 0 0 L†2 D ib ˜
Á ∂ ∂y -i b 0 ˜¯
Ë x

where

1 Ê ∂2 ∂2 ˆ ∂ v- ∂ -
L†2 D = + - b 2 ˜ + u- + - ib w (57)
Re ÁË ∂x 2
∂y 2
¯ ∂ x ∂ y

The key qualitative difference between Eqs. (50) and (52) on the one hand and
either of its counterparts Eqs. (42–43), in case of an one-dimensional background
state, or Eq. (48), in case of a weakly-non-parallel flow, is the coupled discretiza-
tion of the two inhomogeneous spatial directions in Eq. (52); an analogous
observation holds for the corresponding adjoint problems. Rather than a funda-
mental one, this difference is technical: one must abandon well-tested algorithms
of dense-matrix algebra and attempt to solve the resulting numerical problems in
some iterative manner. Some progress has been made recently in both the linear
instability and the linear control aspects of the issue; see Theofilis (2003) for a
review of the former and Jaimoukha and Kasenally (1994) with respect to the
latter issue. Alternatively, one may attempt to reduce the dimensionality of the
flow control problem by projecting the original problem on an (arbitrarily
defined) basis consisting of flow (BiGlobal) eigenmodes and proceed with a
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 109

linear-systems control problem based on this reduced-order model; this approach


was first employed successfully by Hoepffner (2006) and will be discussed fur-
ther shortly.
Rather paradoxically in terms of the associated computing effort, it has been a
non-parallel flow problem which first appeared in the literature of adjoint-based
receptivity and control, namely the influential work of Hill (1992); the work on
the cylinder and the subsequent work on the flat-plate boundary layer (Hill, 1995)
have re-awakend interest in flow control by adjoint methods and firmly estab-
lished the latter as an efficient means of accomplishing this task. It should also be
noted that, in the context of non-parallel background flows, the (numerical effi-
ciency) differences between the adjoint based receptivity and the linear-systems
approach to the flow control problem are even more pronounced; while Hill
(1992) was able to solve numerically the problem of control of the cylinder wake
nearly two decades ago, it was not until recently that the first linear-system
approaches to non-parallel flows have started appearing in the literature (Gallaire
et al., 2004; Marquet et al., 2006). Hill (1992) showed that, besides solution of
the (direct) eigenvalue problem, the essential elements for prediction of the
changes in a flowfield are the associated adjoint eigenvalue problem and an inho-
mogeneous adjoint boundary value problem. The success of his theoretical
approach was based on the fact that there exist two qualitatively different zones
of instability in the cylinder wake, namely a globally unstable followed by a
convectively unstable zone (Huerre and Monkewitz, 1985, 1990). Hill (1992)
presented the direct and adjoint eigenvalue problems in a cylindrical coordinate
system, equivalent to the Cartesian formulation (52–53) and (56–57), and gave
an explicit form for the bilinear concomitant (16). He went on to associate the
small-diameter control cylinder with a feedback control mechanism, modeled as
a momentum point source, and to derive an inhomogeneous adjoint problem, the
solution to which is utilized in order to determine the modification of the global
amplification/damping rate on account of the flow control. Figure 15 summarizes
the results, in terms of the modification (increase) in critical Reynolds number
resulting from the placement of a control cylinder in the wake of the primary
cylinder. The axis of symmetry separates the experimental results of Strykowski
and Sreenivasan (1990), shown in the upper half of the picture, from the theoreti-
cal predictions of Hill (1992), shown in the lower half; excellent agreement may
be seen. Hill also mentioned that the proposed adjoint-based approach is capable
of delivering receptivity results for different kinds of forcing; presumably due to
the large cost of the associated two-dimensional eigenvalue problem computa-
tions, this statement was first corroborated by results in the subsequent work on
the BL (Hill, 1995). In a related study, Giannetti and Luchini (2007) also
employed adjoint based methods to quantify what they called structural instabil-
ity, i.e., perturbations associated with modifications of the direct linear operator.
The maxima of direct/adjoint eigenvector products quantified by these authors
were shown to be associated with regions of maximum flow receptivity. The
receptivity to different sources of flow modification, namely momentum forcing
and mass-injection, is shown in Fig. 16.
Marquet et al. (2006) addressed the issue of control of the global eigenmode of
laminar separation bubble, discovered by Theofilis in an adverse-pressure-gradient
flat-plate BL flow (Theofilis, 2000; Theofilis et al., 2000). Rather than the flat
110 V. THEOFILIS

Fig. 15 Isolines of the shift in critical Reynolds number as a result of the placement
of a control cylinder in the wake, obtained by Hill (1992).

plate, these authors focused on the bubble formed at an engine intake, modeled by
a rounded backward facing step (Fig. 17). The basic state and the direct eigenvalue
problems were solved at a low Reynolds number, at which the basic flow is steady
and laminar. Particular attention was then paid to the long-time optimal perturba-
tion, which is the initial perturbation concentrating energy in the global mode at
large times, via the solution of the adjoint eigenvalue problem associated with an
energy scalar product and the bi-orthogonality conditions (Tumin, 1996, 2003;
Schmid and Henningson, 2001). The leading adjoint global mode results of
Marquet et al. (2006) are shown in Fig. 17; the dominance of the streamwise
component of the disturbance velocity (a result analogous to that of the direct
problem discussed by Theofilis et al. (2000) on the flat plate), was associated with

Fig. 16 Analysis of the structural sensitivity of the cylinder wake by Giannetti and
Luchini (2007), showing receptivity to a) momentum forcing and initial conditions
and to b) mass injection.
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 111

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view

LIVE GRAPH
Click here to view LIVE GRAPH
Click here to view
Fig. 17 Amplitude functions of the leading adjoint eigenmode of rounded-backstep
flow (Marquet et al., 2006).

the energy concentration at the upstream corner of the backstep. Finally, optimal
perturbations were studied, by appropriately defining a cost function analogous to
that introduced by Corbett and Bottaro (2001) in the flat-plate context and making
use of an adjoint initial value problem of the class of (55) in the particular case of
BiGlobal instabilities; the main result was the identification of the mechanism
responsible for amplification of two-dimensional transient perturbations as the
Kelvin–Helmholtz eigenmode of the shear-layer associated with the laminar sepa-
ration bubble.
Finally, work based on adjoint-based instability analysis and optimization on the
one hand and linear-systems control on the other, both applied to essentially non-
parallel flows, has started appearing in the literature. Brion et al. (2007) demonstrated

Fig. 18 Direct and adjoint disturbance vorticity components of a vortex dipole;


shown are the lateral component of the adjoint eigenvector (left side) and the axial
component of the direct eigenvector (right side), both superposed upon the basic flow
streamlines (Brion et al., 2007).
112 V. THEOFILIS

that exciting a counter-rotating laminar vortex dipole with the adjoint eigenmode
of (56) which corresponds to the most unstable (Crow) modal perturbation that
(52) yields, leads to a two- to three-order of magnitude increase in the energy
growth of the leading modal perturbation. In other words, in an aircraft trailing-
vortex context, identification and use of the adjoint Crow mode drastically acceler-
ates the leading modal linear instability of this system, and potentially leads to
substantially faster loss of coherence of the trailing vortex system; Fig. 18 shows
the spatial distribution of both the direct and adjoint Crow modes.
Worth mentioning in the context of linear-systems control of nonparallel flows
are the pioneering efforts of Hoepffner and co-workers (2006) who introduced the
concept of state estimation mentioned earlier (Chevalier et al., 2007) and applied
it to the problem of cavity induced separation (Hoepffner et al., 2007). This flow
is different from the open cavity configuration, the instability and control of which
was respectively studied by Rowley et al. (2002, 2006), in that the dimension of
the former cavity is of the order of the incoming BL thickness and its four corners
are rounded, which significantly facilitates the numerical task of coupled spatial
discretization; while a tensor-product spectral collocation grid suffices for the
former problem, multidomain computations must be performed for the latter, in
order to deal with the corner singularities. In this respect, the configuration of
Hoepffner et al. (2007) has certain geometric similarities with the rounded back-
ward facing step, the instability and control of which was studied by Marquet
et al. (2006). In contrast to the adjoint based optimization approach followed by
the latter authors, building a linear-systems approach for a non-parallel flow
remains a numerically challenging problem (Kim and Bewley, 2007). Hoepffner
et al. (2007) have dealt with this issue by designing a reduced-order model of the
flow prior to the control work, on the premise of the existence of a small subset of
the dynamically-relevant flow eigenmodes. An LQG feedback control scheme
was then constructed on the basis of the reduced-order model, including stochastic
external perturbations. Both the estimation and the control gain matrices were
computed by solutions of algebraic Riccati equations of the class (36) and the
control information was introduced into a direct numerical simulation methodology.
Results obtained include the direct and adjoint eigenvectors, shown in Fig. 19,
as well as demonstrations that the flow-control system constructed is capable of
minimizing the energy of perturbations; this result is shown in Fig. 20, where the
uncontrolled flow (which is subject to modal amplification) is compared with the
flow resulting from application of control after 200 time units; energy growth is
eliminated in the latter case.

E. 2.5- and 3-D: Weakly and Essentially Three-Dimensional Flows


The most general case of three-dimensional background flows, i.e., flows in
which no spatial direction may be treated as homogeneous in the terminology of
Table 1, is the one in which practically no theoretical flow instability (and even less
so theoretical flow control) work exists. The reason is the lack of tools appropriate
for the analysis of such flows and the extreme cost of a full three-dimensional
DNS. This is unfortunate since several interesting applications exist, which could
profit from the development of such tools but do not afford the simplifications of
the background flows made in the previous subsections. Again limiting attention to
ROLE OF INSTABILITY THEORY IN FLOW CONTROL 113

Fig. 19 Direct and adjoint eigenvectors of cavity-induced separated flow (Hoepffner


et al., 2007).

aeronautics applications, examples include flows in three-dimensional (empty or


full bay) cavities, flow around isolated finite-size humps, at engine intakes and
realistic near-field trailing vortex systems, to name but a few.
One recent interesting development in the latter context has been the work of
Broadhurst and Sherwin (2008) (see also Broadhurst et al. (2006) for background
and motivation) who have demonstrated the extensibility of the PSE concept to
three dimensions, whereby two are fully resolved in a coupled manner and the
third is treated as parabolic. This approach is an area of current active research

LIVE GRAPH
Click here to view

Fig. 20 Suppression of energy growth in cavity-induced separated flow (Hoepffner


et al., 2007).
114 V. THEOFILIS

with an aim to complement it by the appropriate adjoint equations in order to


produce a powerful analysis tool for the control of not only complex vortical
flows, but also of weakly three-dimensional fields on non-axisymmetric bodies of
revolution such as elliptic cones in hypersonic flow. Regarding flow control, in the
author’s view, the only realistic hope for near-term theoretically founded results in
this area will come from a combination of reduced-order modeling techniques.
Experience with the latter (Noack et al., 2004; Hoepffner, 2006) seems consis-
tently to suggest that linear instability results are essential ingredients of the
reduced-order-model work. In this respect, research in the areas of reduced-order-
modeling, adjoint based optimization and linear-systems approach to flow control,
on the one hand, and solution of large-scale eigenvalue problems which form
the basis for instability analysis of complex flows on the other, are expected to
mutually benefit each other.

V. Discussion
It is hoped that the short preceding presentation has revealed the current lack
of completeness in our knowledge of instability physics in the complex applica-
tion areas of concern to the present volume. Far from confirming the statement
on maturity of the field of fluid mechanics, the emerging picture is on the one
hand rather promising, as the pioneering theoretical flow control work in the
previous decade has shown, and on the other hand full of remaining challenges,
especially in terms of issues associated with the analysis and control of multi-
dimensional flow applications, abundantly present in real-world configurations.
The field of (theoretical) flow control via control of flow instabilities is already
large and growing, but may still be considered to be at its early development
stages, and it would be unwise to risk predicting the directions it may take in
future. Nevertheless, the current state of the art is quite some distance away from
becoming off-the-shelf technology, even for the simplest of one-dimensional
background flows. This is even less the case in the main driver of flow control
applications, namely flows over complex geometries, and less well explored flow
regimes such as hypersonic flow. Further theoretical flow control contributions
are welcome, as are solutions to practical problems arising in multi-dimensional
flows; reduced-order-modeling techniques (including theoretical foundation
thereof) as well as techniques for the solution of (massive) eigenvalue problems
pertinent to instability and control of inhomogeneous flows would be pivotal in
introducing the techniques demonstrated on model flows to the real world.

Acknowledgments
This work was sponsored by the Air Force Office of Scientific Research and
the European Office of Aerospace Research and Development, under contracts
F49620-03-1-0295, FA8655-03-1-3059 and FA8655-06-1-3066, monitored by
Thomas Beutner (now at DARPA), Rhett Jefferies (AFOSR), John Schmisseur
(AFOSR) and Surya Surampudi (EOARD). Interactions with colleagues in the
framework of these projects, as well as those within the EU-STREP project
“Far-Wake” AST4-CT-2005-012238, are kindly appreciated.
Chapter 5

Dynamic and Closed-Loop Control

Clarence W. Rowley*
Princeton University, Princeton, New Jersey
and
Belinda A. Batten†
Oregon State University, Corvalis, Oregon

I. Introduction
In this chapter we present techniques for feedback control, focusing on those
aspects of control theory most relevant for flow control applications.
Feedback control has been used for centuries to regulate engineered systems.
In the 17th century, Cornelius Drebbel invented one of the earliest devices to use
feedback, an incubator that used a damper controlled by a thermometer to main-
tain a constant temperature (Franklin et al., 2005). Feedback devices became
more prevalent during the industrial revolution, and James Watt’s speed governor
for his steam engine has become one of the most well known early examples of
feedback regulators (Åström and Murray, 2008).
In recent decades, the use of feedback has proliferated and, in particular, it has
advanced the aerospace industry in critical ways. The ability of feedback control-
lers to modify the dynamics of a system, particularly to stabilize systems that are
naturally unstable, enabled the design of aircraft such as the F-16, whose longitu-
dinal dynamics are unstable, dramatically increasing the performance envelope.
This ability of feedback systems to modify a system’s natural dynamics is dis-
cussed further in Secs. II.A and IV.E.
Only recently have closed-loop controllers been used in flow control applications.
Our objective here is to outline the main tools of control theory relevant to these
applications, and discuss the principal advantages and disadvantages of feedback

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Associate Professor, Mechanical and Aerospace Engineering. Associate Fellow AIAA.
†Professor, Mechanical Engineering. Member AIAA.

115
116 C. W. ROWLEY AND B. A. BATTEN

control, relative to the more common open-loop flow control strategies. We also aim
to provide a bridge to the controls literature, using notation and examples that we
hope are more accessible to a reader familiar with fluid mechanics, but possibly not
with a background in controls. Since it is obviously not possible to give a detailed
explanation of all of control theory in a single chapter, here we merely touch on
highlights, and refer the reader to the literature for further details.

II. Architectures
Within this section, we describe the basic architecture of a closed-loop system,
discussing reasons for introducing feedback, as opposed to strategies that do not
use sensors, or use sensors in an open-loop fashion. The general architecture of a
closed-loop control system is shown in Fig. 1, where the key idea is that informa-
tion from the plant is sensed, and used in computing the signal to send to the actua-
tor. To describe the behavior of these systems, we will require mathematical models
of the system to be controlled, called the plant, as shown in Fig. 1. We then design
our controller based on a model of the plant, and assumptions about sensors, actua-
tors, and noise or disturbances entering into the system. We will see that some
properties such as controllability and observability are fixed by the architecture of
the closed-loop system, in particular what we choose as sensors and actuators.

A. Fundamental Principles of Feedback


Why should one use feedback at all? What are the advantages (and disadvan-
tages) of feedback control over other control architectures? In practice, there are
many tradeoffs (including weight, added complexity, reliability of sensors, etc.),
but here we emphasize two fundamental principles of feedback.

1. Modify Dynamics
Feedback can modify the natural dynamics of a system. For instance, using
feedback, one can improve the damping of an underdamped system or stabilize an
unstable operating condition, such as balancing an inverted pendulum. Open-loop

f Plant y
P

Controller
C

Fig. 1 Typical block diagram for closed-loop control. Here, P denotes the plant, the
system to be controlled, and C denotes the controller, which we design. Sensors and
actuators are denoted y and f, respectively, and d denotes external disturbances. The
minus sign in front of f is conventional, for negative feedback.
DYNAMIC AND CLOSED-LOOP CONTROL 117

or feed-forward approaches cannot do this.* For flow control applications, an


example is keeping a laminar flow stable beyond its usual transition point. We will
give specific examples of modifying dynamics through feedback in Sec. IV.E.

2. Reduce Sensitivity
Feedback can also reduce sensitivity to external disturbances or to changing
parameters in the system itself (the plant). For instance, an automobile with a
cruise control that senses the current speed can maintain the set speed in the
presence of disturbances such as hills, headwinds, etc. In this example, feedback
also compensates for changes in the system itself, such as changing weight of the
vehicle as the number of passengers changes. If instead of using feedback, a
lookup table were used to select the appropriate throttle setting based only on
the desired speed, such an open-loop system would not be able to compensate
either for disturbances or changes in the vehicle itself.

B. Models of Multi-Input, Multi-Output (MIMO) Systems


A common way to represent a system is using a state-space model, which is a
system of first-order ordinary differential equations (ODEs) or maps. If time is
continuous, then a general state-space system is written as

q = F (q, f , d), q(0) = q 0


y = G(q, f , d) (1)

where q(t) is the state, f(t) is the control, d(t) is a disturbance, and y(t) is the
measured output. We often assume that F is a linear function of q, f, and d, for
instance obtained by linearization about an equilibrium solution (for fluids, a
steady solution of Navier–Stokes). It can also be convenient to define a controlled
output z(t), which we choose so that the objective of the controller is to keep z(t)
as small as possible. For instance, z(t) is often a vector containing weighted
versions of the sensed output y and the input f, because often the objective is to
keep the outputs small, while using little actuator power.
Many of the tools for control design are valid only for linear systems, so the
form of Eq. (1) is often too general to be useful. Instead, we often work with linear
approximations. For instance, if the system depends linearly on the actuator f, and
if the sensor y and controlled output z are also linear functions of the state q, then
Eq. (1) may be written

q = Aq + N (q) + Bf + Dd, q(0) = q 0 (2)

y = C1q + D11 f + D12 d (3)

z = C2 q + D21 f + D22 d. (4)

*This is true at the linear level—if nonlinearities are present, however, then open-loop forcing can

stabilize an unstable operating point, as demonstrated by the classic problem of a vertically forced
pendulum (Verhulst, 1996).
118 C. W. ROWLEY AND B. A. BATTEN

The term Aq is the linear part of the function F above, and N(q) the nonlinear
part (for details of how these are obtained, see Sec. VI). The term C1q repre-
sents the part of the state that is sensed and C2q the part of the system that is
desired to be controlled. The matrices D, D12, and D22 are various disturbance
input operators to the state, measured output, and controlled output; the distur-
bance terms may have known physics associated with them, or may be consid-
ered to be noise. Hence, Eq. (2) provides a model for the dynamics (physics)
of the state, and the influence of the control through the actuators on those
dynamics, Eq. (3) provides a model of the measured output provided by the
sensors, and Eq. (4) a model of the control objective.
When the plant is modeled by a system of ODEs, the state will be an element
of a vector space, e.g., q(t) 僆 Rn for fixed t. In a case where the plant is mod-
eled by a system of partial differential equations (PDEs), e.g., as with the
Navier–Stokes equations, the same notation can be used. In this context, the
notation q(t) is taken to mean q(t, ·), where t is a fixed time and the other
independent variables (such as position) vary. This interpretation of the nota-
tion leads to q(t) representing a “snapshot” of the state space at fixed time. If W
denotes the spatial domain in which the state exists, then q(t) lies in a state
space that is a function space, such as L2(W) (the space of functions that are
square integrable, in the Lebesgue sense, over the domain W). The results in
this chapter will be presented for systems modeled by ODEs, although there
are analogs for nearly every concept defined for systems of PDEs. We do this
for two primary reasons. First, the ideas will be applied in practice to compu-
tational models that are ODE approximations to the fluid systems, and this will
give the reader an idea of how to compute controllers. Second, the PDE analogs
require an understanding of functional analysis and are outside the scope of
this book. We point out that the fundamental issue that arises in applying these
concepts to ODE models that approximate PDE models is one of convergence,
that is, whether or not quantities such as control laws, computed for a dis-
cretized system, converge to the corresponding quantities for the original PDE.
We note that it is not enough to require that the ODE model of the system, e.g.,
one computed from a computational fluid dynamics package, converge to the
PDE model. We refer the interested reader to Curtain and Zwart (1995) for
further discussion on control of PDE systems.
Throughout much of this chapter, we will restrict our attention to the simplified
linear systems

q (t ) = Aq(t ) + Bf (t ) q(0) = q 0
(5)
y ( t ) = C q( t )

or

q (t ) = Aq(t ) + Bf (t ) q(0) = q 0
y(t ) = C1q(t ) (6)
z(t ) = C2 q(t ).
DYNAMIC AND CLOSED-LOOP CONTROL 119

The first of these forms is useful when including sensed measurements with the
state, and the second is useful when the control objective involves particular states
given by z, different from the sensed quantities. Note that in perhaps the most
common case in which feedback is used, controlling about an equilibrium (such
as a steady laminar solution), the linearization (5) is always a good approximation
of the full nonlinear system (1), as long as we are sufficiently close to the
equilibrium.* Since the objective of control is usually to keep the system close to
this equilibrium, the linear approximation is often useful even for systems which
naturally have strong nonlinearities.
The fundamental assumption is typically made that the plant is approximately
the same as the model. We note that the model is never perfect and typically con-
tains inaccuracies of various forms: parametric uncertainty, unmodeled dynamics,
nonlinearities, and perhaps purposely neglected (or truncated) dynamics. These
factors motivate the use of a robust controller as given by MinMax or other H •
controllers that will be discussed in Sec. IV.D.

C. Controllability, Observability
Given an input–output system (e.g., a physical system with sensors and actua-
tors), two characteristics that are often analyzed are controllability and obser-
vability. For a mathematical model of the form (1), controllability describes the
ability of the actuator f to influence the state q. Recall that, for a fluid, the state is
typically all flow variables everywhere in space at a particular time, so controlla-
bility addresses the effect of the actuator on the entire flow. Conversely, observ-
ability describes the ability to reconstruct the full state q from available sensor
measurements y. For instance, in a fluid flow, often we cannot measure the entire
flowfield directly. The sensor measurements y might consist of a few pressure
measurements on the surface of a wing, and we may be interested in reconstructing
the flow everywhere in the vicinity of the wing. Observability describes whether
this is possible (assuming our model of the system is correct).
The following discussion assumes we have a linear model of the form (5). Such
a system is said to be controllable on the time interval [0, T] if, given any initial
state q0 and final state q f, there exists a control f that will steer from the initial
state to the final state.
Controllability is a stringent requirement: for a fluid system, if the state q con-
sists of flow variables at all points in space, this definition means that, for a system
to be controllable, one must be able to drive all flow variables to any desired values
in an arbitrarily short time, using the actuator. Models of fluid systems are there-
fore often not controllable, but if the uncontrollable modes are not important for
the phenomena of interest, they may be removed from the model, for instance
using methods discussed in Sec. V.
Fortunately, controllability is a property that is easy to test: a necessary and
sufficient condition is that the controllability matrix

C = [B AB A2 B  An-1 B] (7)

*And as long as the system is not degenerate: in particular, A cannot have eigenvalues on the imagi-

nary axis.
120 C. W. ROWLEY AND B. A. BATTEN

have full rank (rank = n). As before, n is the dimension of the state space, so q is
a vector of length n and A is an n × n matrix.
The system (5) is said to be observable if for all initial times t0 the state
q(t0) can be determined from the output y(t) defined over a finite time interval
t 僆 [t0, t1]. The important point here is that we do not simply consider the output
at a single time instant, but rather watch the output over a time interval, so that at
least a short time history is available. This idea lets us reconstruct the full state
(e.g., the full flow field) from what is typically a much smaller number of sensor
measurements.
For the system to be observable, it is necessary and sufficient that the obser-
vability matrix
È ˘
C Í ˙
Í ˙
CA Í
Í
˙
˙
Í ˙
O = CA2 Í
Í
˙
˙
(8)
Í ˙
Í
Í
 ˙
˙
Í n - 1˙
CA
ÍÎ ˙˚

have full rank. For a derivation of this condition, and of the controllability test (7),
see introductory controls texts such as Bélanger (1995).
Although in theory the rank conditions of the controllability and observability
matrices are easy to check, these can be poorly conditioned computational opera-
tions to perform. A better way to determine these properties is via the controllabil-
ity and observability Gramians.
An alternative test for controllability of system (5) involves checking the rank of
an n × n symmetric matrix called the time-T controllability Gramian, defined by
T *
Wc (T ) = Ú e At BB*e A t dt . (9)
0

This matrix Wc(T) is invertible if and only if the system (5) is controllable.
Similarly, an alternative test for observability on the interval [0, T] is to check
invertibility of the observability Gramian, defined by
T *
*
Wo (T ) = Ú e A t C Ce At dt . (10)
0

(In these formulas, * denotes the adjoint of a linear operator, which is simply trans-
pose for matrices, or complex-conjugate transpose for complex matrices.) One can
show that the ranks of the Gramians in Eqs. (9, 10) do not depend on T, and for a
stable system we can take T Æ •. In this case, the (infinite-time) Gramians Wc and
Wo may be computed as the unique solutions to the Lyapunov equations

AWc + Wc A* + BB* = 0 (11)

A*Wo + Wo A + C *C = 0. (12)
DYNAMIC AND CLOSED-LOOP CONTROL 121

These are linear equations to be solved for Wc and Wo, and one can show that as
long as A is stable (all eigenvalues in the open left half of the complex plane), they
are uniquely solvable, and the solutions are always symmetric, positive-semidefinite
matrices. These Gramians are more useful than the controllability and observ-
ability matrices defined in Eqs. (7) and (8) for studying systems that are close to
uncontrollable or unobservable (i.e., one or more eigenvalues of Wc or Wo are almost
zero, but not exactly zero). One can, in fact, construct examples which are arbitrarily
close to being uncontrollable (an eigenvalue of Wc is arbitrarily close to zero), but
for which the controllability matrix (7) is the identity matrix (Dullerud and Paganini,
1999)! While the Gramians are harder to compute, they contain more information
about the degree of controllability: the eigenvectors corresponding to the largest
eigenvalues can be considered the “most controllable” and “most observable” direc-
tions, in ways that can be made precise. For derivations and further details, as well
as more tests for controllability and observability, and the related concepts of
stabilizability and detectability, see Dullerud and Paganini (1999). For further dis-
cussion and algorithms for computation, see also Datta (2004).

D. State Space vs Frequency Domain


In the frequency domain, the models in Sec. II.B are represented by transfer
functions, to be discussed in more detail in Sec. III.B. By taking the Laplace
transform of the system in Eq. (5), we obtain the representation of the system

y (s) = C (sI - A)-1 B f (s) (13)

where G(s) = C(sI - A)-1B is the transfer function from input to sensed output. For
the system in Eq. (6), we have the additional transfer function H(s) = C2(sI - A)-1B
as the transfer function from input f to controlled output z. Many control concepts
were developed in the frequency domain and have analogs in the state space. The
type of approach that is used often depends on the goal of applying the control
design. For example, if one is trying to control behavior that has a more obvious
interpretation in the time domain, the state space may be the natural way to con-
sider control design. On the other hand, if certain frequency ranges are of interest
in control design, a frequency domain approach may be desired. Many references
consider both approaches and for more information we refer the reader to standard
controls texts (Anderson and Moore, 1990; Åström and Murray, 2008; Dorato
et al., 2000; Franklin et al., 2005; Kwakernaak and Sivan, 1972; Skogestad and
Postlethwaite, 1996; Zhou et al., 1996).

III. Classical Closed-Loop Control


In this section, we explore some features of feedback control from a classical
perspective. Traditionally, classical control refers to techniques that are in the
frequency domain (as opposed to state-space representations, which are in the
time domain), and often are valid only for linear, single-input, single-output sys-
tems. Thus, in this section, we assume that the input f and output y are scalars,
denoted f and y, respectively. The corresponding methods are often graphical, as
they were developed before digital computers made matrix computations relatively
122 C. W. ROWLEY AND B. A. BATTEN

easy, and they involve using tools such as Bode plots, Nyquist plots, and root-locus
diagrams to predict behavior of a closed-loop system.
We begin by describing the most common type of classical controller, proportional-
integral-derivative (PID) feedback. We then discuss more generally the notion of
a transfer function, and Bode and Nyquist methods for predicting closed-loop
behavior (e.g., stability or tracking performance) from open-loop characteristics
of a system. For further information on the topics in this section, we refer the
reader to standard texts (Åström and Murray, 2008; Franklin et al., 2005).

A. PID Feedback
By far the most common type of controller used in practice is proportional-
integral-derivative (PID) feedback. In this case, in the feedback loop shown in
Fig. 1, the controller is chosen such that

• d
f (t ) = - Kp y(t ) - K i Ú y(t ) dt - K d y( t ) (14)
0 dt

where Kp, Ki, and Kd are proportional, integral, and derivative gains, respectively.
In general, for a first-order system, PI control (in which Kd = 0) is generally effec-
tive, whereas for a second-order system, PD control (in which Ki = 0) or full PID
control is usually appropriate.
To see the effect of PI and PD feedback on these systems, consider a first-order
system of the form
.
y + ay = f (15)
where f is the input (actuator), y is the output (sensor), and a is a parameter, and
suppose that the control objective is to alter the dynamics of Eq. (15). Without
feedback ( f = 0), the system has a pole at -a, or a dynamic response of e-at.
We may use feedback to alter the position of this pole. Choosing PI feedback
f = -Kpy - Ki Ú y dt, the closed-loop system becomes


y + (a + K p ) y + K i y = 0. (16)

The closed-loop system is now second order and, with this feedback, the poles are
the roots of

s 2 + (a + Kp ) s + K i = 0. (17)

Clearly, by appropriate choice of Kp and Ki, we may place the two poles anywhere
we desire in the complex plane, since we have complete freedom over both coef-
ficients in Eq. (17).
Next, consider a second-order system, a spring-mass system with equations of
motion

my + by + ky = f (18)


DYNAMIC AND CLOSED-LOOP CONTROL 123

where m is the mass, b is the damping constant, and k is the spring constant. With
.
PD feedback, u = -Kpy - Kd y, the closed-loop system becomes

my + (b + Kd ) y + (k + Kp ) y = 0 (19)

with closed-loop poles satisfying

ms 2 + (b + Kd ) s + k + Kp = 0. (20)

Again, we may place the poles anywhere we desire, since we have freedom to
choose both of the coefficients in Eq. (20). For tracking problems (where the
control objective is for the output to follow a desired reference), often integral
feedback is used to drive the steady-state error to zero. However, introducing
integral feedback into second-order systems can make the closed loop unstable for
high gains, so the techniques in the latter part of this chapter need to be used to
ensure stability. Also, note that in practice if the gains Kp and especially Kd are
increased too much, sensor noise becomes a serious problem. For more informa-
tion and examples of PID control, we refer the reader to conventional controls
texts (Åström and Murray, 2008; Franklin et al., 2005).

B. Transfer Functions
Linear systems obey the principle of superposition, and hence many useful
tools are available for their analysis. One of these tools that is ubiquitous in clas-
sical control is the Laplace transform. The Laplace transform of a function of time
y(t) is a new function ỹ(s) of a variable s 僆 C, defined as

y (s) = L{y(t )} = Ú e -st y(t ) dt (21)
0

defined for values of s for which the integral converges. For properties of the Laplace
transform, we refer the reader to standard texts (Åström and Murray, 2008; Marsden
and Hoffman, 1998), but the most useful property for our purposes is that

Ï dy ¸
L Ì ˝ = sy (s) - y(0) (22)
Ó dt ˛

Hence, the Laplace transform converts differential equations into algebraic


equations.
For a plant with input u and output y and dynamics given by

y + a1 y + a0 y = b1 f + b0 f
a2  (23)
.
taking the Laplace transform, with zero initial conditions (y(0) = y(0) = f(0) = 0),
gives

(a2 s 2 + a1 s + a0 ) y = (b1 s + b0 ) f (24)


124 C. W. ROWLEY AND B. A. BATTEN

or

y (s ) b1 s + b0
= = P(s) (25)
u(s ) a2 s 2 + a1 s + a0

where P(s) is called the transfer function from f to y. Thus, the transfer function
relates the Laplace transform of the input to the Laplace transform of the output.
It is often defined as the Laplace transform of the impulse response (a response to
an impulsive input, f(t) = d(t)), since the Laplace transform of the Dirac delta
function is 1. For a state-space representation as in Eqs. (2) and (3), the transfer
function is given by Eq. (13).
The transfer function has a direct physical interpretation in terms of the frequency
response. If a linear system is forced by a sinusoidal input at a particular frequency,
once transients decay the output will also be sinusoidal, at the same frequency but
with a change in amplitude and a phase shift. That is,

f (t ) = sin(w t ) fi y(t ) = A(w )sin(w t + j (w )). (26)

The frequency response determines how these amplitudes A(w) and phase shifts
j(w) vary with frequency. These may be obtained directly from the transfer func-
tion. For any frequency w, P(iw) is a complex number, and one can show that

A(w ) = | P(iw ) |,
j (w ) = –P(iw ) (27)
______
where for a complex number z = x + iy, |z| = ÷x2 + y2 and –z = tan-1 y/x. The
frequency response is often depicted by a Bode plot, which is a plot of the magni-
tude and phase as a function of frequency (on a log scale), as shown in Fig. 2.
The amplitude is normally plotted on a log scale or in terms of decibels where
A(dB) = 20 log10 A.

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view
a) Bode Diagram
Gm = 9.54 dB (at 3.32 rad/sec), Pm = 44.5 deg (at 1.84 rad/sec)
b) Nyquist Diagram
20 2.5

0 2
Magnitude (dB)

−20
1.5
−40
1
−60
Imaginary Axis

0.5
−80 1/GM
0
−100
0 PM
–0.5
−45
–1
Phase (deg)

−90
−135 –1.5

−180 –2
−225
–2.5
−270 –2 –1 0 1 2 3 4
10−1 100 101 102 Real Axis
Frequency (rad/sec)

Fig. 2 Plots for the loop gain P(s)C(s) for the example system (29), showing the gain
and phase margins (denoted GM and PM): a) Bode and b) Nyquist.
DYNAMIC AND CLOSED-LOOP CONTROL 125

C. Closed-Loop Stability
Bode plots tell us a great deal about feedback systems, including information
about the performance of the closed-loop system over different ranges of frequen-
cies and even the stability of the closed loop system and how close it is to becom-
ing unstable.
For the system in Fig. 1, suppose that the transfer function from d to y is W(s),
the transfer function from f to y is P(s), and the transfer function of the controller
is C(s), so that the system has the form shown in Fig. 3. Then the closed-loop
transfer function from disturbance d to the output y is

y (s ) W (s)
 = . (28)
d ( s ) 1 + P(s )C (s )

Closed-loop poles therefore occur where 1 + P(s)C(s) = 0 so in order to investi-


gate stability, we study properties of the loop gain G(s) = P(s)C(s).
One typically determines stability using the Nyquist criterion, which uses
another type of plot called a Nyquist plot. The Nyquist plot of a transfer function
G(s) is a plot of G(iw) in the complex plane (Im G(iw) vs Re G(iw)), as w varies
from -• to •. _____
Note that since the coefficients in the transfer function G(s) are
real, G(-iw) = G(iw), so the Nyquist plot is symmetric about the real axis. The
Nyquist stability criterion then states that the number of unstable (right-half-plane)
closed-loop poles of Eq. (28) equals the number of unstable (right-half-plane)
open-loop poles of G(s), plus the number of times the Nyquist plot encircles the
-1 point in a clockwise direction. [This criterion follows from a result in complex
variables called Cauchy’s principle of the argument (Franklin et al., 2005)].
One can therefore determine the stability of the closed loop just by looking at the
Nyquist plot and counting the number of encirclements of the –1 point. For most
systems that arise in practice, this criterion is also straightforward to interpret in
terms of a Bode plot. We illustrate this through an example in the following section.

D. Gain and Phase Margins and Robustness


Suppose that we have designed a controller that stabilizes a given system. One
may then ask how much leeway we have before the feedback system becomes
unstable. For instance, suppose our control objective is to stabilize a cylinder wake
at a Reynolds number at which the natural flow exhibits periodic shedding, and that

d Plant
W(s)

f + + y
P(s)

C(s)

Fig. 3 More detailed block diagram of feedback loop in Fig. 1.


126 C. W. ROWLEY AND B. A. BATTEN

we have designed a feedback controller that stabilizes the steady laminar solution.
How much can we adjust the gain of the amplifier driving the actuator, before the
closed-loop system loses stability and shedding reappears? Similarly, how much
phase lag can we tolerate, for instance in the form of a time delay, before closed-
loop stability is lost? Gain and phase margins address these questions.
As a concrete example, consider the system in Fig. 3, with transfer functions

1
P(s) = , C (s ) = 20 (29)
(s + 1)(s + 2)(s + 3)

for which the Nyquist and Bode plots of G(s) = P(s)C(s) are shown in Fig. 2. We
see that the Nyquist plot does not encircle the -1 point so, for this value of C(s),
the closed-loop system is stable by the Nyquist stability criterion. However, if the
gain were increased enough, the Nyquist plot would encircle the -1 point twice,
indicating two unstable closed-loop poles. The amount by which we may increase
the gain before instability is called the gain margin, and is indicated GM in the
figure. The gain margin may be determined from the Bode plot as well, as also
shown in the figure, by noting the value of the gain where the phase crosses -180 deg.
This value is one measure of how close the closed-loop system is to instability: a
small gain margin (close to 1) indicates the system is close to instability.
Another important quantity is the phase margin, also shown in Fig. 2 (denoted
PM). The phase margin is specifies how much the phase at the crossover frequency
(where the magnitude is 1) differs from -180 deg. Clearly, if the phase equals
-180 deg at crossover, the Nyquist plot intersects the -1 point, and the system is
on the verge of instability. Thus, the phase margin is another measure of how close
the system is to instability.
Both of these quantities relate to the overall robustness of the closed-loop
system, or how sensitive the closed-loop behavior is to uncertainties in the model
of the plant. Of course, our mathematical descriptions of the form (29) are only
approximations of the true dynamics, and these margins give some measure of
how much imprecision we may tolerate before the closed-loop system becomes
unstable. See Doyle et al. (1992) for a more quantitative (but still introductory)
treatment of these robustness ideas.

E. Sensitivity Function and Fundamental Limitations


One of the main principles of feedback discussed in Sec. II.A is that feedback
can reduce a system’s sensitivity to disturbances. For example, without control
(C(s) = 0 in Fig. 3), the transfer function from the disturbance d to the output y is
W(s). With feedback, however, the transfer function from disturbance to output is
given by Eq. (28). That is, the transfer function is modified by the amount

1
S (s) = (30)
1 + G(s)

called the sensitivity function, where G(s) = P(s)C(s) is the loop gain, as before.
Thus, for frequencies w where |S(iw)| < 1, then feedback acts to reduce the effect
DYNAMIC AND CLOSED-LOOP CONTROL 127

of disturbances; but if there are frequencies for which |S(iw)| > 1, then feedback
has an adverse effect, amplifying disturbances more than without control.
The magnitude of the sensitivity function is easy to see graphically from the
Nyquist plot of G(s), as indicated in Fig. 4. At each frequency w, the length of the
line segment connecting -1 to the point G(iw) is 1/|S(iw)|. Thus, for frequencies
for which G(iw) lies outside the unit circle centered at the -1 point, disturbances
are attenuated, relative to the open-loop system. However, we see that there are
some frequencies for which the Nyquist plot enters this circle, for which
|S(iw)| > 1, indicating that disturbances are amplified, relative to open-loop.
The example plotted in Fig. 4 used a particularly simple controller. One might
wonder if it is possible to design a more sophisticated controller such that distur-
bances are attenuated at all frequencies. Unfortunately, one can show that under
very mild restrictions it is not possible to do this for any linear controller. This fact
is a consequence of Bode’s integral formula, also called the area rule (or the
“no-free-lunch theorem”), which states that if the loop gain has relative degree ≥2
(i.e., G(s) has at least two more poles than zeros), then the following formula
holds:

Ú0
log10 | S (iw ) | dw = p(log10 e) Â Re p
j
j (31)

where pj are the unstable (right-half-plane) poles of G(s) (for a proof, see Doyle
et al., 1992). The limitations imposed by this formula are illustrated in Fig. 5,

Nyquist diagram for G(s)


2.5

2
LIVE GRAPH
Click here to view
1.5
ΩS(iω)Ω< 1
1
Attenuation
Imaginary Axis

0.5 ΩS(iω)Ω> 1
Amplification
0

–0.5

–1

–1.5 1/ΩS(iω)Ω G(iω)

–2

–2.5
–2 –1 0 1 2 3 4
Real Axis

Fig. 4 Nyquist plot of the loop gain G(s) = P(s)C(s) for the system (29). For frequen-
cies for which G(iw) enters the unit circle centered about the −1 point, disturbances
are amplified and, for frequencies for which G(s) lies outside this circle, disturbances
are attenuated relative to open-loop.
128 C. W. ROWLEY AND B. A. BATTEN

10

LIVE GRAPH
Click here to view
5 ΩS(iω)Ω

+
Magnitude (dB)


–5

–10

–15
1 2 3 4 5 6 7 8 9 10
Frequency (rad/sec)

Fig. 5 Magnitude of S(iw), illustrating the area rule (31): for this system, the area of
attenuation (denoted - ) must equal the area of amplification (denoted +), no matter
what controller C(s) is used.

which shows |S(iw)| for the G(s) used in our example. Here, one sees that the area
of attenuation, for which log|S(iw)| < 0, is balanced by an equal area of amplifica-
tion, for which log|S(iw)| > 0. These areas must, in fact, be equal in this log-linear
plot. If our plant P(s) had unstable poles, then the area of amplification must be
even greater than the area of amplification, as required by Eq. (31).
While these performance limitations are unavoidable, typically in physical situa-
tions, it is not a severe limitation, since the area of amplification may be spread over a
large frequency range with only negligible amplification. However, in some examples,
particularly systems with narrow bandwidth actuators in which these regions must be
squeezed into a narrow range of frequencies, this result can lead to large peaks in
S(iw), leading to strongly amplified disturbances for the closed-loop system. These
principles explain peak-splitting phenomena that have been observed in several flow-
control settings, including combustion instabilities (Banaszuk et al., 2006b) and cav-
ity oscillations (Rowley and Williams, 2006; Rowley et al., 2006). This peaking
phenomenon is further exacerbated by the presence of time delays, as described in the
context of a combustor control problem in Cohen and Banaszuk (2003).

IV. State-Space Design


We discuss the fundamental approaches to feedback control design in the time
domain in the subsections below. We start with the most fundamental problems
and add more practical aspects. By the end of this section the reader should have
an overview of some of the most common approaches to linear control design for
state-space models.
DYNAMIC AND CLOSED-LOOP CONTROL 129

A. Full-State Feedback: Linear Quadratic Regulator Problem


In this section, we assume that we have knowledge of all states (i.e., C1 is the
identity matrix) and can use that information in a feedback control design. Although
this is typically infeasible, it is a starting point for more practical control designs.
The linear quadratic regulator (LQR) problem is stated as follows: find the control
f(t) to minimize the quadratic objective function


J (f ) = Ú (qT (t )Qq(t ) + f T (t ) Rf (t )) dt (32)
0

subject to the state dynamics

q (t ) = Aq(t ) + Bf (t ), q(0) = q 0 . (33)

The matrices Q and R are state and control weighting operators and may be chosen
to obtain desired properties of the closed-loop system. In particular, when there is
a controlled output as in Eq. (4), defining Q = C2TC2 places the controlled output in
the objective function. It is necessary that Q be non-negative definite and that R
be positive definite in order for the LQR problem to have a solution. In addition,
(A, B) must be controllable (as defined in Sec. II.C). This problem is called linear
because the dynamic constraints are linear, and quadratic since the objective
function is quadratic. The controller is called a regulator because the optimal
control will drive the state to zero.
The solution to this problem can be obtained by applying the necessary condi-
tions of the calculus of variations. One finds that under the above assumptions, the
solution to this problem is given by the feedback control

f(t) = -Kq(t), K = R-1BT P

where the matrix K is called the feedback gain matrix, and the symmetric matrix
P is the solution of the algebraic Riccati equation

A*P + P A - P BR -1BT P + Q = 0. (34)

This is a quadratic matrix equation, and has many solutions P, but only one
positive-definite solution, which is the one we desire. Equations of this type
may be solved easily, for instance using the Matlab commands are or lqr.
Often, one wishes to drive the state to some desired state qd. In that case, a
variant of this problem called a tracking problem can be solved (Dorato et al.,
2000).

B. Observers for State Estimation


Typically, one does not have the full-state available, but rather only a (possibly
noisy) sensor measurement of part of the state, as described in Sec. II. For these
problems, in order to implement a full-state feedback law as described in the
130 C. W. ROWLEY AND B. A. BATTEN

previous section, one needs an estimator (or filter or observer) that provides an
estimate of the state based upon sensed measurements. The estimator is another
dynamic system, identical to the state equation (5), but with an added correction
term based on the sensor measurements:

q c (t ) = Aq c (t ) + Bf (t ) + L (y(t ) - Cq c (t )), q c (0) = q c 0 . (35)

Without the correction term, this equation is identical to (5) and hence our esti-
mate qc(t) should match the actual state q(t), as long as we know the initial value
q(0), and as long as our model (5) perfectly describes the actual system. However,
in practice, we do not have access to a perfect model or to an initial value of the
entire state, so we add the correction term L(y - Cqc), and choose weights L so
that the state estimate qc(t) approaches the actual state q(t) as time goes on. In
particular, defining the estimation error e = qc - q, and combining Eqs. (5) and
(35), we have
.
e(t) = (A - LC)e(t) (36)

so this error converges to zero as long as the eigenvalues of A - LC are in the open
left half plane. One can show (see, for instance, Bélanger, 1995) that as long as the
pair (A, C) is observable, the gains L may be chosen to place the eigenvalues of
A - LC anywhere in the complex plane (the pole-placement theorem). Hence, as
long as the system is observable, the weights L can always be chosen so that the
estimate converges to the actual state.
The weights L can also be chosen in an optimal manner, which balances the
relative importance of sensor noise and process noise. If one has a noisy or
inaccurate sensor, then the measurements y should not be trusted as much, and
the corresponding weights L should be smaller. Process noise consists of actual
disturbances to the system [e.g., d in Eq. (2)], so if these are large, one will need
larger sensor corrections to account for these disturbances, and in general the
entries in L should be larger. The optimal compromise between sensor noise and
process noise is achieved by the Kalman filter. There are many types of Kalman
filters, for continuous-time, discrete-time, time-varying, and nonlinear systems.
Here, we discuss only the simplest form, for linear time-invariant systems
(Bélanger, 1995) and, for more complex cases, we refer the reader to Gelb (1974),
Stengel (1994) or other references (Dorato et al., 2000; Kwakernaak and Sivan,
1972; Willems, 2004).
We first augment our model with terms that describe the sensor noise n and
disturbances d:

q = Aq + Bf + Dd
(37)
y = Cq + n.

We additionally assume that both n and d are zero-mean, Gaussian white noise,
with covariances given by

E(ddT ) = Qe, E(nnT ) = Re (38)


DYNAMIC AND CLOSED-LOOP CONTROL 131

where E(·) denotes the expected value. The optimal estimator (that minimizes the
covariance of the error e = qc - q in steady state) is given by L = PC TR -1
e ,where
P is the unique positive-definite solution of the Riccati equation

AP + PAT - PCTR -1 T
e CP + DQe D = 0. (39)

This optimal estimation problem is, in a precise sense, a dual of the LQR problem
discussed in Sec. IV.A and may also be solved easily, using Matlab commands
are or lqe.

C. Observer-Based Feedback
Once we have an estimate qc of the state, for instance from the optimal observer
described above, we can use this estimate in conjunction with the state feedback
controllers described in Sec. IV.A, using the feedback law

f(t) = -Kqc(t). (40)

Combining the above equation with the observer dynamics (35), the resulting
feedback controller depends only on the sensed output y, and is called a Linear
Quadradic Gaussian (LQG) regulator.
One might well question whether this procedure should work at all: if the state-
feedback controller was designed assuming knowledge of the full state q, and then
used with an approximation of the state qc, is the resulting closed-loop system
guaranteed to be stable? The answer, at least for linear systems, is yes, and this
major result is known as the separation principle: if the full-state feedback law
f = -Kq stabilizes the system (e.g., A - BK is stable), and the observer (35) con-
verges to the actual state (e.g., A – LC is stable), then the observer based feedback
law (40) stabilizes the full system.
One can see this result as follows. Combining the controller given by Eq. (40)
with the estimator dynamics (35) and the state dynamics (33), one obtains the
overall closed-loop system

È q (t ) ˘ È A - BK ˘ È q(t ) ˘
Í ˙ = Í ˙ Í q (t ) ˙ . (41)
Îq c(t )˚ Î LC A - BK - LC ˚ Î c ˚

Changing to the variables (q, e), where as before, e = qc - q, Eq. (41) becomes

Èq (t )˘ È A - BK - BK ˘ Èq(t )˘
Í ˙ = Í ˙Í ˙. (42)
Î e (t ) ˚ Î 0 A - LC ˚ Î e(t ) ˚

Since the matrix on the right-hand side is block upper triangular, its eigenvalues
are the union of the eigenvalues of A - BK and those of A - LC. These are always
stable (eigenvalues in the open left-half plane) if the state-feedback law is stable
and the observer dynamics are stable, and so the separation principle holds.
This remarkable result demonstrates that one is justified in designing the state
feedback gains K separately from the observer gains L.
132 C. W. ROWLEY AND B. A. BATTEN

D. Robust Controllers: MinMax Control


If the state equation has a disturbance term present as
.
q(t) = Aq(t) + Bf(t) + Dd(t), q(0) = q0 (43)

there is a related control design that can be applied, the MinMax design (Başar
and Bernhard, 1995; Datta, 2004; Rhee and Speyer, 1989). The objective function
to be solved is to find

Min Max J ( f , d) = Ú (qT (t ) Q q(t ) + f T(t ) R f (t ) - g 2 dT(t ) d (t )) dt (44)
f d 0

subject to Eq. (43). The solution has the same feedback form as in the LQR prob-
lem, but this time, the Riccati matrix P is the solution to

AT P + P A - P[BR-1BT - q2DDT]P + Q = 0 (45)

where q = 1/g. The MinMax control is more robust to disturbances than is LQR;
the larger the value of q, the more robust it is. There is a limit to the size of q and
if it is chosen as large as possible, the resulting controller is a different kind of
optimal controller called the H• controller which is robust in a sense that can be
made precise, but can be overly conservative (see Datta (2004) and Mustafa and
Glover (1990) for details).
The MinMax estimator can be determined through a similar augmentation of
Eq. (39). Specifically, the algebraic Riccati equation

AP + PAT - P[CTC - q 2Q]P + DDT = 0 (46)

is solved for P, and q is taken to be as large as possible. In particular, P and P are


the minimal positive semi-definite solutions to the algebraic Riccati equations
(the matrix [I - q 2Pq Pq] must also be positive definite).
If one defines K = R-1BT P as above and

L = [ I - q 2 PP]-1 PC T (47)

Ac = A - BK - LC + q 2 DDT P (48)

then the MinMax controller for the system (43) is given by

f(t) = -r -1BT Pzc(t) = -Kzc(t). (49)

Observe that for q = 0 the resulting controller is the LQG (i.e., Kalman filter)
controller.
We make the following note regarding this theory as applied to PDE systems.
The theory for LQG and MinMax control design exists for such systems and
requires functional analysis. Application of the methods requires convergent
approximation schemes to obtain systems of ODEs for computations. Although
this can be done in the obvious way—approximate the PDE by finite differences,
DYNAMIC AND CLOSED-LOOP CONTROL 133

finite elements, etc.—convergence of the state equation is not enough to ensure


convergence of the controller. This can be seen when one considers the algebraic
Riccati equations and notes that for the PDE, the term AT is replaced by the adjoint
of A, A*. A numerical approximation scheme that converges for A might not con-
verge for A*. This should not dissuade the interested reader from applying these
techniques to fluid control problems. It is noted, however, as an issue that one
must consider, and there are many results in the literature on this topic (Burns and
King, 1998).
At this stage, the full-order LQG control is impractical for implementation; as
for many applications of interest, the state estimate is quite large, making the
controller intractable for real-time computation. There are many approaches to
obtaining low-order controllers for both infinite and finite dimensional systems.
Some involve reducing the state-space model before computing the controllers,
while others involve computing a control for the full-order system—known to
converge to the controller for the PDE systems—and then reducing the controller
in some way. In Sec. V, we discuss some reduction techniques that have been
explored for low-order control design.

E. Examples
Our first example is a nonlinear system of ordinary differential equations given by

x1
x1 (t ) = - + x2 - x2 x12 + x22 + d
R
2x
x 2 (t ) = - 2 + x1 x12 + x22 + d
R

where R is a parameter to be specified. We will assume that d is a small distur-


bance. Here, the system matrices are given by

È ˘
È-1/R 1 ˘ Í
Í
- x2 x12 + x22 ˙˙
A =Í N ( x) =
-2 /R ˙˚
Í ˙
Î 0 Í
ÎÍ
x1 x12 + x22 ˙
˚˙

È1˘ È0˘
D=Í˙ d (t ) = e B = Í ˙.
Î1˚ Î1˚

Depending upon the choice of R and d, the system has various equilibria. We will
choose R = 5 and look at phase plots of the uncontrolled system, as shown in
Fig. 6. There are five equilibria: three stable, including the origin, and two unstable.
When introducing a small disturbance of d = 10-3, there is little visible change in
the behavior of the system, as shown in the subfigure on the right. The green lines
indicate a band of stable trajectories that are drawn to the stable equilibrium at
the origin—about which the system was linearized.
If we increase to R = 6 and perform the same experiment, we find that the
inclusion of the disturbance greatly changes the system, annihilating two of the
equilibria, and dramatically decreasing the band of stable trajectories (Fig. 7).
134 C. W. ROWLEY AND B. A. BATTEN
LIVE GRAPH LIVE GRAPH
a) Click here to view b) Click here to view
1 1

0.5 0.5

x2
x2 0 0

−0.5 −0.5

−1 −1
−1 −0.5 0 0.5 1 −1 −0.5 0 0.5 1
x1 x1

Fig. 6 Phase plots of the uncontrolled system for R = 5: a) no disturbance and


b) d = 10−3.

To apply the control methodology discussed above, we linearize the system


around an equilibrium to design a control. We choose the origin about which to
linearize, and design LQR and MinMax controls. The phase plot of the nonlinear
system with LQR feedback control included is shown in Fig. 8 (the plot is visually
identical with and without a disturbance). Note that the LQR control expands the
region that is attracted to the origin, but the sinks at the upper right and lower left
regions of the plot still survive, and trajectories outside of the s-shaped region
converge to these.
When the MinMax control is applied to the system, a marked change in the
phase plot occurs, as shown in Fig. 9. Now, the origin is globally attracting:
the sinks at upper left and lower right disappear, and all trajectories are drawn to
the origin. Note that this behavior is not necessarily an anomaly of this example.
LIVE GRAPH
Click here to view

0.5

x2 0

−0.5

−1

−1 −0.5 0 0.5 1
x1

Fig. 7 Phase plot of the uncontrolled system for R = 6, d = 10−3, showing the annihi-
lation of two of the equilibria seen in Fig. 6.
DYNAMIC AND CLOSED-LOOP CONTROL 135

1
LIVE GRAPH
Click here to view
0.5

x2 0

–0.5

–1

–1 –0.5 0 0.5 1
x1

Fig. 8 Phase plot of the LQR controlled system for R = 5, d = 0 or d = 10−3. The goal
is to stabilize the origin, and the region of attraction is the s-shaped strip between the
two gray lines.

That is, linear feedback control can be surprisingly effective at fundamentally


altering nonlinear systems, and stabilizing their equilibria.
Our second example involves a PDE model with limited sensed measurements
and actuation. This is a model for the vibrations in a cable-mass system. The cable
is fixed at the left end and attached to a vertically vibrating mass at the right as
shown in Fig. 10. The mass is then suspended by a spring that has a nonlinear
stiffness term. We assume that the only available measurements are the position
and velocity of the mass, and that a force is applied at the mass to control the

1
LIVE GRAPH
Click here to view
0.5

x2 0

–0.5

–1

–1 –0.5 0 0.5 1
x1

Fig. 9 Phase plot of the MinMax system for R = 5, d = 0 or d = 10−3, showing that all
trajectories eventually reach the origin.
136 C. W. ROWLEY AND B. A. BATTEN

Fig. 10 Cable-mass system.

structure. In addition, there is a periodic disturbance applied at the mass of the


form d(t) = a cos(wt). The equations for this system are given below:

∂2 ∂ È ∂ ∂2 ˘
r 2
w(t , s) = Ít w(t , s) + g w(t , s)˙
∂t ∂s Î ∂s ∂t ∂s ˚
2 2
∂ È ∂ ∂ ˘
m 2 w(t , ) = - Ít w(t , ) + g w(t , )˙ - a1w(t , )
∂t Î ∂ s ∂ t ∂ s ˚
-a 3 [ w(t , )]3 + d (t ) + u(t )

w(t, 0) = 0 (50)

w(0, s ) = w0 (s ), w(0, s ) = w1 (s ). (51)
∂t

We refer the interested reader to Burns and King (1998) for discussion about the
approximation scheme applied to the problem for computational purposes.
For this problem, we linearize the system about the origin, and design LQG and
MinMax controllers with estimators based on the sensed measurements. In Fig. 11,
the mass position with respect to time is shown for the two control designs. We see
a greater attenuation of the disturbance with MinMax control.
In Fig. 12, we show the phase plot of the controlled mass as compared to the
uncontrolled mass. Again, note the greater disturbance attenuation under MinMax
control.
It might not be too surprising to note that MinMax control does a better job of
controlling the mass—the part of the system where the measurements are taken
and where actuation occurs. Fig. 13 shows behavior of the mid-cable, and we see
the great attenuation of the MinMax control once again.

V. Model Reduction
The methods of analysis and control synthesis discussed in the previous two
sections rely on knowledge of a model of the system, as a system of ODEs,
DYNAMIC AND CLOSED-LOOP CONTROL 137
LIVE GRAPH
Click here to view
a) LQG compensator b) MinMax compensator
2 2

1 1
Mass position

Mass position
0 0

–1 –1

–2 –2
0 50 100 150 200 0 50 100 150 200
Seconds Seconds LIVE GRAPH
Click here to view
Fig. 11 Mass position: a) LQG controlled and b) MinMax controlled.

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view
a) Open loop vs. LQG b) Open loop vs. MinMax
10 10

5 5
Mass velocity

Mass velocity

0 0

–5 –5

–10 –10
–4 –2 0 2 4 –4 –2 0 2 4
Mass position Mass position

Fig. 12 Mass position: a) LQG control and b) MinMax control, comparing uncon-
trolled case (. . .) with controlled (—).

Fig. 13 Mid-cable phase plot: a) LQG control and b) MinMax control.


138 C. W. ROWLEY AND B. A. BATTEN

represented either as a transfer function or in state-space form (2). However, the


equations governing fluid motion are not ODEs—they are PDEs, and even if the
resulting PDEs are discretized and written in the form (2), the number of states
will be very large, equal to the number of gridpoints × flow variables, which typi-
cally exceeds 106. While it is possible to compute optimal controllers for the
Navier–Stokes equations even in full three-dimensional geometries (Bewley et al.,
2001; Högberg et al., 2003), in order to be used in practice these approaches would
require solving multiple Navier–Stokes simulations faster than real time, which is
not possible with today’s computing power.
However, for many fluids problems, one does not need to control every detail
of a turbulent flow, and it is sufficient to control the dominant instabilities or
the large-scale coherent structures. For these simpler control problems, it often
suffices to use a simplified numerical model, rather than the full Navier–Stokes
equations. The process of model reduction involves replacing a large, high-fidelity
model with a smaller, more computationally tractable model that closely approxi-
mates the original dynamics in some sense.
The techniques described in this section give an overview of some methods of
model reduction that are useful for fluid applications. However, we emphasize
that obtaining reduced-order models that accurately capture a flow’s behavior
even when actuation is introduced remains a serious challenge and is a topic of
ongoing research. In many turbulent flows, there may not even exist a truly low-
dimensional description of the flow that will suffice for closed-loop control. Thus,
the techniques discussed here are most likely to be successful when one has reason
to believe that a low-dimensional description would suffice: for instance, flows
dominated by a particular instability, or flows that exhibit a limit cycle behavior
(see Sec. VI.B for a discussion of limit cycles).
In this section, we first discuss Galerkin projection onto basis functions deter-
mined by Proper Orthogonal Decomposition (POD), a method used by Aubry
et al. (1988), Holmes et al. (1996), Lumley (1970), Sirovich (1987) and others.
Next, we discuss balanced truncation (Moore, 1981), a method which has been
applied to fluids comparatively recently (Cortelezzi et al., 1998; Lee et al., 2001;
Rowley, 2005). There are many other methods for model reduction, such as Hankel
norm reduction (Obinata and Anderson, 2000), which has even better performance
guarantees than balanced truncation. However, most other methods scale as n3,
and so are not computationally tractable for large fluids simulations where n ~ 106.
See Antoulas et al. (2001) for an overview of several different model reduction
procedures for large-scale systems.

A. Galerkin Projection
We begin with dynamics that are defined on a high-dimensional space, say
q 僆 Rn for large n:
.
q = F(q, f), q 僆 Rn (52)
where as before, f denotes an input, for instance from an actuator. Typically,
trajectories of the system (52) do not explore the whole phase space Rn, so we may
DYNAMIC AND CLOSED-LOOP CONTROL 139

wish to approximate solutions by projections onto some lower-dimensional subspace


S Ã Rn. Denoting the orthogonal projection onto this subspace by PS: Rn Æ S Ã Rn,
Galerkin projection defines dynamics on S simply by projecting the following:
.
q = PSF(q, f), q 僆 S. (53)

For instance, if S is spanned by some basis functions {ϕ1, . . . , ϕr}, then we may
write
r

q(t) = Âa (t)ϕ
j=1
j j (54)

and the reduced equations (52) have the form

·ϕ , ϕ Ò a. (t) = ·ϕ , F(q(t), f(t))Ò


k
j k k j (55)

which are ODEs for the coefficients ak(t). Here, ··, ·Ò denotes an inner product,
which is just dot product for vectors, but other inner products may be used. If the
basis functions are orthogonal, then the matrix formed by ·ϕj, ϕkÒ is the identity,
.
so the left-hand side of Eq. (55) is simply aj.
Note that this procedure involves two choices: the choice of the subspace S, and
the choice of the inner product. For instance, if v and w are in Rn (regarded as
column vectors), and if ·v, wÒ = vTw denotes the standard inner product, then
given any symmetric, positive-definite matrix Q, another inner product is

·v, wÒQ = vTQw. (56)


Different choices of inner product lead to quite different dynamics, and a suitable
choice can guarantee certain useful properties. For instance, if one chooses an
“energy-based” inner product (such that the “energy” ||q|| Q2 = ·q, qÒQ is conserved
or decreases), then the reduced-order model (55) is guaranteed to preserve stabil-
ity of an equilibrium point at q = 0, something that is not guaranteed otherwise
(Rowley et al., 2004). We will see that balanced truncation corresponds to choos-
ing a particularly useful inner product that also satisfies this property.

B. Proper Orthogonal Decomposition


POD is one method of determining basis functions jj to use for Galerkin
projection as described in the previous section. In particular, POD determines modes
that are optimal for capturing the most energetic features in a given data set.
Because we are often interested in PDEs, let us assume we are given a set of
data q(x, t), a vector of flow variables q, as a function of space x and time t (here,
x may be a vector). We wish to approximate q as a sum
n

q̃(x, t) = Âa (t)ϕ (x)


j=1
j j (57)
140 C. W. ROWLEY AND B. A. BATTEN

where the functions ϕj(x) are fixed, vector-valued basis functions (modes), and
we will in addition assume these modes are orthonormal. Note that because the
modes ϕj in Eq. (57) are fixed, the flow at any time t is completely specified by
the coefficients aj(t). These coefficients may be computed from our data q(x, t)
using orthonormality:

Ú
aj (t) = ·q(x, t), ϕj (x)Ò = qT(x, t), ϕj (x) dx
W
(58)

where W is the spatial domain. We want to find modes that minimize the average
~ and the original data q.
error between the projected data q
One can show—e.g., using a variational argument (Holmes et al., 1996)—that
the optimal modes are solutions to the infinite-dimensional eigenvalue problem
____________
Ú
Rϕ (x) = lϕ (x), Rϕ (x) = q(x, t) qT(x¢, t) ϕj (x¢) dx¢
W
(59)

Here, R is an operator that takes a function of space ϕ(x) and creates another
function of space, given by the right-hand side of Eq. (59), and the overbar repre-
sents an appropriate average (e.g., time average). The eigenvalues l represent the
“energy captured” by each POD mode.
In finite dimensions, the integral in Eq. (59) becomes a sum, and the POD modes
may be found by standard eigenvalue solvers or by singular value decomposition.
For instance, suppose the data is a scalar variable q in a single spatial dimension x,
given at certain points in space and time as q(xj, tk), with j = 1, . . . , n, k = 1, . . . , m.
Then the POD modes are the eigenvectors of the real symmetric matrix
m
1
Rij =
m  q( x , t )q( x , t )
k =1
i k j k (60)

This is an n × n eigenvalue problem, but if the number of snapshots m is smaller


than the number of gridpoints n, it is more efficient to compute the POD methods
using the method of snapshots (Sirovich, 1987), described below.
First, form the data matrix Ajk = q(xj, tk), whose columns are the snapshots:

È q( x1 , t1 ) q( x1 , t2 )  q( x1 , t m ) ˘
Í q( x , t ) q( x , t )  q( x , t ) ˙
A=Í 2 1 2 2 2 m ˙
(61)
Í    ˙
Í ˙
Îq( xn , t1 ) q( xn , t2 )  q( xn , t m ) ˚

Then the following are equivalent:


1) The POD modes are the eigenvectors of the n × n matrix R = (1/m) AAT (the
direct method).
2) The POD modes are given by jj = Acj where cj 僆 Rm are eigenvectors of the
m × m matrix M = (1/m) ATA [the method of snapshots (Sirovich, 1987)].
DYNAMIC AND CLOSED-LOOP CONTROL 141

3) The POD modes are left singular vectors of A. That is, writing a singular value
decomposition A = USV *, the POD modes are the columns of U.
For problems in two or three spatial dimensions, one can stack values of q at all
gridpoints into a single column vector in each column of A in matrix (61). (In this
case the order in which the gridpoints are stacked is not important.) If q is a vector,
then one may compute separate sets of modes for each flow variable or a single set
of vector-valued modes. The latter approach is always advantageous for incom-
pressible flows, since if the individual snapshots satisfy the continuity equation
div q(x, t) = 0, then the individual vector valued modes will also satisfy continuity
(Holmes et al., 1996). Vector-valued modes have also been shown to behave better
in examples in compressible flow as well, although here one must be careful to
define a suitable inner product involving both velocities and thermodynamic
variables (Rowley et al., 2004).
For more in-depth tutorials of POD, see Chatterjee (2000) or Holmes et al.
(1996). There are many other extensions to this basic approach as well, including
methods for computing modes from incomplete data (Everson and Sirovich,
1995), traveling POD modes (Rowley and Marsden, 2000), scaling POD modes
for self-similar solutions (Aronson et al., 2001; Beyn and Thümmler, 2004;
Rowley et al., 2003), and shift modes (Noack et al., 2003) for better capturing
transient behavior.

C. Balanced Truncation
The POD/Galerkin approach has been successful for a large number of flow
problems. However, the method is often quite fragile: the models depend unpre-
dictably on the number of modes kept, and often a large number of modes is
required to capture qualitatively reasonable dynamics. One potential problem
with the philosophy of POD/Galerkin models is that the choice of modes is based
on retaining the most energetic features, and low-energy phenomena (such as
acoustic waves) may be important to the overall dynamics. Balanced truncation is
another (related) method of model reduction, widely used in the control theory
community. It is not based solely on energetic importance and seems to produce
more reliable models than POD/Galerkin in examples (Antoulas et al., 2001; Ilak
and Rowley, 2006; Rowley, 2005). The method also has provable error bounds
that are close to the minimum possible for any reduced-order model of a given
order (Dullerud and Paganini, 1999).
Balanced truncation applies to linear input-output systems, for instance of the
form Eq. (5). The philosophy of balanced truncation is to truncate the modes that are
least controllable (that are the least excited by inputs u), and are least observable (that
have the smallest effect on future outputs y). The degree of controllability and observ-
ability is defined quantitatively in terms of the Gramians defined in Eqs. (9) and (10).
In the original coordinates these two criteria may be contradictory, so balanced trun-
cation proceeds by first transforming to coordinates in which the controllability and
observability Gramians are equal and diagonal, via a balancing transformation. It is
always possible to find such a change of coordinates as long as the original system is
both controllable and observable. In the transformed coordinates one then keeps only
the first few modes which are the most controllable and observable.
142 C. W. ROWLEY AND B. A. BATTEN

For algorithms for computing balancing transformations, see Dullerud and


Paganini (1999) or Datta (2004). The standard approach involves first solving the
Lyapunov equations (11–12), and then computing a transformation that simulta-
neously diagonalizes the Gramians. Because the Gramians are full (non-sparse)
n × n matrices, this approach becomes computationally intractable for large n, as
occurs in many fluids problems. For this reason, the snapshot-based approach in
Lall et al. (2002) can be useful, and an algorithm for computing the balancing
transformation directly from snapshots without first computing the Gramians is
also available (Ilak and Rowley, 2008; Rowley, 2005).

VI. Nonlinear Systems


While the tools we have studied throughout most of this chapter have addressed
linear systems, virtually all real systems are nonlinear. Fortunately, as we have
seen in the examples in Sec. IV.E, linear techniques often work surprisingly well
even for nonlinear systems. However, it is helpful to understand characteristics of
nonlinear systems, both to understand the limitations of linear control methods
and to better understand the behavior of real systems.
Here, we give only an elementary overview of the most basic aspects of non-
linear systems. For a more detailed introduction, see Hirsch et al. (2004) and, for
a more advanced treatment, see Wiggins (1990), Khalil (1996), and Guckenheimer
and Holmes (2002).

A. Multiple Equilibria and Linearization


Nonlinear systems may be written in the general form
.
q = F(q, f, µ, t) (62)

where q(t) and f(t) are the state vector and input, as in Sec. IV, and µ 僆 Rp is a
vector of parameters (for instance, Reynolds number). To apply the control meth-
ods discussed in the first part of this chapter, we must rewrite Eq. (62) in the form
(5). To do this, we first identify equilibrium points of the system (62), which are
.
points where q = 0, or F(q, f, µ, t) = 0. While linear systems typically have only
one equilibrium point (at q = 0), nonlinear systems may have multiple equilibria,
as illustrated by the example in Sec. IV.E.
To simplify notation, we will assume that the system does not depend explicitly
on time and for the moment we will suppress the dependence on the parameter µ,
.
writing q = F(q, f). Suppose that (q*, f *) is an equilibrium point of (62), so that
F(q*, f ) = 0. Expanding F in a Taylor series about (q*, f *), we have
*

∂F * *
F (q* + d q, f * + d f ) = F (q*, f * ) + (q , f ) ◊d q
∂q
∂F * *
+ (q , f ) ◊d f + Higher order terms
∂f (63)
DYNAMIC AND CLOSED-LOOP CONTROL 143

where

Ê ∂F1 ∂F1 ∂F1 ˆ



Á ∂x1 ∂x 2 ∂x n ˜
Á ˜
Á ∂F2 ∂F2 ∂F2 ˜
∂F * * 
(q , f ) ∫ Á ∂x1 ∂x 2 ∂x n ˜ (64)
∂q Á ˜
Á    ˜
Á ˜
Á ∂Fn ∂Fn

∂Fn ˜
ÁË ∂x ∂x 2 ∂xn ˜¯ q = q* ,f = f *
1

is the Jacobian matrix of F, evaluated at the point (q*, f *) [also called the deriva-
tive at (q*, f *), and often denoted DF(q*, f *)], and ∂F/∂f is the matrix of partial
derivatives with respect to the components of f. Note that we evaluate these matri-
ces of partial derivatives at the equilibrium point (q*, f *), so these are just constant
matrices.
In linearization, we neglect the higher-order terms, so letting

∂F * * ∂F * *
A= (q , f ), B= (q , f ) (65)
∂q ∂f

.
the equation q = F(q, f) becomes

q * + d q = F (q*, f ) + A ◊d q + B ◊d f
*

Now, since (q*, f *) is an equilibrium point, F(q*, f *) = 0, and the equation


becomes

d q = A ◊d q + B ◊d f (66)

which is of the form presented in system (5).


If one wishes to include nonlinear terms, as in Eq. (2), one defines

N (d q, d f ) = F (q* + d q, f * + d f ) - A d q - B d f . (67)

Note that in general, N in Eq. (2) may depend on d f as well as d q, although the
control-affine form given in Eq. (2) is quite common in practice.
Similarly, if the output y is a nonlinear function y = G(q, f ), the output equation
may be similarly linearized, writing y = y* + ∂y, with y*= G(q*, f *), and

d y = Cd q + Dd f (68)
144 C. W. ROWLEY AND B. A. BATTEN

where

∂G * * ∂G * *
C= (q , f ), D= (q , f ). (69)
∂q ∂f

B. Periodic Orbits and Poincaré Maps


Nonlinear systems that arise in fluid mechanics often exhibit periodic orbits,
which are simply periodic solutions q(t) of form (62). When these periodic orbits
are stable (or stable in reverse time), they are called limit cycles. Examples include
vortex shedding in the wakes of bluff bodies (Noack et al., 2003), or oscillations
in aeroelasticity problems (Dowell, 1980).
A common tool for studying periodic orbits is the Poincaré map, which is
defined as follows: if the overall phase space has dimension n (that is, q 僆 Rn),
one takes an n - 1-dimensional cross section S of phase space, and defines a map
F : S Æ S where F(q0) is found by evolving the system (62) forward in time with
q(0) = q0, until q(t) once again intersects S. The point F(q0) is then defined as this
point of intersection, as illustrated in Fig. 14.
The system dynamics may then be represented by the discrete-time system

q k + 1 = F (q k ) (70)
.
and periodic orbits in the original system q = F(q) are equilibrium points (where
F(q) = q) of the discrete-time system (70). Approaches are also available for
including an input in such systems, and linearizing about the periodic orbit (see,
for instance, Farhood et al., 2005).

C. Simple Bifurcations
Often, when parameters are varied, one observes qualitative changes in pheno-
mena. In fluid mechanics these qualitative changes are referred to as transitions;

Φ(x0)

x0

Fig. 14 Three-dimensional phase space (n = 3), showing the Poincaré section S and
the corresponding Poincaré map F.
DYNAMIC AND CLOSED-LOOP CONTROL 145

in dynamical systems these are referred to as bifurcations. Here, we give a brief


overview of some common bifurcations. For more information, see standard texts
in dynamical systems (Guckenheimer and Holmes, 2002; Verhulst, 1996; Wiggins,
1990).
In a precise sense, a bifurcation occurs for the differential equation
q = F (q, m ) (71)

or the map
(72)
qk+1 = F(qk, m)

for a parameter value m for which the system is not structurally stable. We will not
define structural stability in a precise sense here (see Guckenheimer and Holmes
(2002) for a definition), but loosely speaking, a system is structurally stable if its
phase portrait remains qualitatively unchanged when the system itself [e.g., the
right-hand side of Eq. (71)] is perturbed. Note that this concept is different from
stability of an equilibrium point (or periodic orbit), in which one considers pertur-
bations in the initial conditions, rather than the system itself.
Two different categories of bifurcations can arise in dynamical systems. Local
bifurcations of equilibria are the simplest to analyze, and occur when equilibrium
points change stability type. An example from fluid mechanics is a laminar flow
losing stability when the Reynolds number increases. When this occurs, new
phenomena can appear, such as new equilibria or new periodic orbits. For instance,
when the Reynolds number is increased for the flow around a cylinder, the initially
stable steady flow (an equilibrium point) transitions to a Karman vortex street (a
periodic orbit). These local bifurcations can be studied by analyzing the behavior
of the system near the equilibrium point of the differential equation (71) or map
(72). Another category of bifurcation is global bifurcations, which involve quali-
tative changes in the phase portrait, without equilibria changing stability type. For
instance, periodic orbits may appear or disappear and invariant manifolds may
change the way they intersect with one another. These are usually more difficult
to detect and analyze.
Here, we will discuss only the simplest types of local bifurcations, those arising
generically when there is a single parameter µ 僆 R (codimension-1 bifurcations).
In order for a bifurcation to occur at the point (q*, m*), two conditions must
be met:
1) The linearization (∂F/∂q)(q*, m*) must have an eigenvalue on the imaginary
axis (or, for maps, on the unit circle).
2) The eigenvalue must cross the imaginary axis (or unit circle) with nonzero
speed.
For differential equations, when a real eigenvalue crosses the imaginary axis,
several different types of bifurcations can occur, including saddle-node, tran-
scritical, and pitchfork bifurcations. These can be distinguished by checking
degeneracy conditions (conditions on the higher derivatives of F; see Guckenheimer
and Holmes, 1983). When a pair of complex eigenvalues crosses the imaginary
axis, a Hopf bifurcation occurs, and in this case, a one-parameter family of periodic
146 C. W. ROWLEY AND B. A. BATTEN

orbits is generated in the neighborhood of the bifurcation point. All of these types
of bifurcation occur for maps as well. In addition, another type of bifurcation
occurs for maps when an eigenvalue crosses the unit circle at the point -1: a flip,
or period-doubling bifurcation. This type of bifurcation is common in fluids (for
instance, arising as bifurcations of a Poincaré map), and an illustration of this type
is shown in Fig. 15. Note that after the period doubling bifurcation occurs, a
periodic orbit of the original period still exists, but is unstable.

D. Characterizing Nonlinear Oscillations


Oscillations occur in a wide variety of fluid systems and, if one wants to control
these oscillations, it is important to understand the fundamental mechanisms that
produce them. One common mechanism is an equilibrium point (steady solution of
Navier-Stokes) becoming unstable for a particular value of a parameter and pro-
ducing a stable limit cycle through a Hopf bifurcation, as described in the previous
section. Limit cycles are commonly observed in aeroelasticity problems (Dowell,
1980), and occur in cavity flows (Rockwell and Naudascher, 1978; Rowley et al.,
2002), cylinder wakes (Noack et al., 2003), and many other types of problem.
Another common mechanism is fundamentally linear in nature: an equilibrium
may be stable but lightly damped (with eigenvalues close to the imaginary axis),
leading to strong resonant frequencies. If such a system is forced by noise, it can
exhibit narrow-band oscillations as well, in much the same way that a tuning fork

a)

p Σ
Im

p
x1 Re
x
–1 1

Time t
b)
Σ
a Im
b
p a
x1 b Re
x
–1 1

Time t

Fig. 15 Period-doubling (flip) bifurcation. The plots at the left show a time history of
one of the state variables a) before and b) after the bifurcation; the center column
shows the phase plane, and the Poincaré map before and after bifurcation; the right
column shows the location of the corresponding eigenvalue of the Poincaré map cross-
ing the unit circle.
DYNAMIC AND CLOSED-LOOP CONTROL 147

would continue to ring at a single frequency if it were continually forced at random


frequencies. For fluid applications, possible sources of this forcing include acous-
tic excitation, wall roughness, or boundary-layer turbulence. This mechanism of
oscillations in fluids has been discovered in a variety of applications as well,
including combustion instabilities (Banaszuk et al., 2006b), flutter in turbo-
machinery (Rey et al., 2003), and cavity flows (Rowley et al., 2006).
These two mechanisms of oscillation suggest fundamentally different types of
control strategies. In the first scenario, possible control objectives are stabilization
of an unstable equilibrium point or reduction in amplitude of a limit cycle. In the
second scenario, the equilibrium point is already stable and, to reduce the amplitude
of oscillations, the control strategy should focus on attenuation of disturbances.
It is often difficult to distinguish between these two types of oscillations purely
from spectra: both are characterized by sharp resonant peaks. However, there are
methods for distinguishing these based purely on measurements (Mezic and
Banaszuk, 2004). One first identifies a peak in the spectrum of a time series, and
passes the data through a narrow-band filter about the frequency of the peak. One
then plots the probability density function (PDF) of the bandpass-filtered data. If
the system possesses a limit cycle, more time will be spent at extrema of the limit
cycle, so the PDF will exhibit two peaks, near the extrema; conversely, if the
system is lightly damped and driven by noise, then more time will be spent near
the average value and the PDF will exhibit a single peak about zero. In this man-
ner, one can distinguish between these two mechanisms of oscillation and deter-
mine the most appropriate control strategy.

E. Methods for Control of Nonlinear Systems


The discussions of control design tools in Sec. IV have been restricted to appli-
cations to linear systems. Although that may seem overly restrictive, that is the
starting point for many real control systems. A basic approach to developing
nonlinear controllers is to take the nominal design from an LQG or MinMax
estimator-based control and to augment it with a nonlinear term. This approach is
referred to as forming an extended filter. This extension could be performed for a
series of set points, obtaining a family of controls. These controls can then be
applied when the system is near the various operating conditions; such an approach
is called gain scheduling.
Other control design methodologies are specifically tailored to nonlinear sys-
tems and for an introduction to these see Khalil (1996). Many of these techniques
apply to restricted classes of systems (e.g., Hamiltonian or Lagrangian systems or
systems with particular types of nonlinearities), and so far have not been widely
used in the flow control community, where models are typically too large or messy
for these techniques to apply, but they may become more widely used as the
models for flow control problems improve.
1) Receding horizon control. One approach for nonlinear systems involves
solving an optimal control problem, for instance minimizing a cost function simi-
lar to Eq. (32), but using the full nonlinear dynamics. For a nonlinear system, one
typically uses a gradient-based search to find a value of f(t) that gives a local
minimum of the cost function for a particular value of the initial state q(0). This
yields an open-loop control law valid only for this specific initial condition.
148 C. W. ROWLEY AND B. A. BATTEN

In practice, one typically evaluates the cost function (32) over a finite time hori-
zon, and then recomputes the optimal open-loop control f(t) as often as possible,
using an updated version of the initial state, in order to account for disturbances
or model uncertainties. This approach, in which the nonlinear optimal control
solution is computed and recomputed in real time, is called receding horizon
control, and is an effective but computationally intensive approach. For more
information on this approach, see Stengel (1994) or Åström and Murray (2008).
2) Adaptive control and extremum seeking. Another type of nonlinear
control design that can be applied to linear design to augment the performance
and robustness of the controller is adaptive control. A strategy that has been
particularly successful in recent experiments (Becker et al., 2006; Henning and
King, 2005; King et al., 2004), especially for situations requiring optimal tuning
of a largely open-loop strategy, is extremum seeking. This is an approach where
one measures a quantity one wants to maximize or minimize (e.g., minimize
drag), and adjusts the value of a control parameter to extremize this quantity. The
tuning is performed by superimposing small variations on the control parameter
and slowly adjusting the mean value of the control parameter depending on
whether the quantity to be extremized is in phase or out of phase with the varia-
tions of the control parameter. This is an old technique, dating back to the 1950s
(Morosanov, 1957; Ostrovskii, 1957), but it is only recently that stability of these
techniques has been proven for certain classes of systems (Krstic and Wang,
2000). This approach has the advantage that it is completely model free, and so
tends to be very robust, though it is not appropriate when control is needed at
timescales similar to those of the natural plant dynamics.
There are many good references on other adaptive control strategies, and we
refer the interested reader to Haykin (1996).

F. Summary
In this chapter, we have given an overview of many topics from control theory,
emphasizing the topics that we believe are the most relevant for design of feedback
controllers for flow control applications. In particular, the techniques described in
this chapter are useful primarily for understanding closed-loop control systems, as
defined in Chapter 3, Sec. III. While the majority of the flow control work to date
has involved open-loop control, closed-loop control is becoming increasingly
common, and the ideas in this chapter (for instance, Secs. II.A and III.E) discuss
the potential benefits and limitations of feedback control. The next chapter (in
particular, Sec. III.B) also discusses sensor requirements specific to closed-loop
control designs.
Later chapters of this book contain several examples in which closed-loop
control has been used in practice. For instance, Chapter 9, Sec. 5 discusses closed-
loop control for turbomachinery applications, and Chapter 10 discusses a number
of applications to the control of combustion instabilities.
Chapter 6

Actuators and Sensors

Louis Cattafesta* and Mark Sheplak†


University of Florida, Gainesville, Florida

I. Introduction
Any active flow control system necessarily includes actuators and, for the case
of systems employing feedback control, sensors. The modeling, design, fabrica-
tion, and testing of these transducers can be complicated, expensive, and time
consuming, often requiring numerous design cycles with less than satisfactory
results. The purpose of this chapter is to review recent progress in popular actua-
tors and microelectromechanical systems (MEMS) unsteady pressure and shear-
stress sensors for flow control applications. The chapter is broken into two parts;
a discussion of actuators precedes that of sensors. As with any engineering system,
a design cannot proceed without a statement of often conflicting technical require-
ments. Therefore each major section begins with a brief discussion of basic termi-
nology and design specifications, followed by their specific interplay with potential
flow control objectives.
In the case of actuators, there are numerous types and space does not allow a
comprehensive discussion of all of these. Therefore only the most popular or
promising are discussed, beginning with some historical perspective. A section is
also included on actuators potentially suitable for high-speed flows. In all cases,
we attempt to summarize key advantages and disadvantages of each, along with
unresolved research issues. Suggestions for metrics to consider when using actua-
tors in active flow control applications are also provided.
In the case of sensor development, a discipline in and of itself, we restrict
our attention to wall-mounted dynamic sensors, which are important because
of the practical requirement that control of real flows requires (ideally) non-
intrusive dynamic sensing at the bounding walls of the flow domain. Attention

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*
Professor, Interdisciplinary Microsystems Group, Department of Mechanical and Aerospace
Engineering. Associate Fellow AIAA.
†Associate Fellow AIAA.

149
150 L. CATTAFESTA AND M. SHEPLAK

is focused on MEMS unsteady surface pressure and wall shear stress sensors
because of their ability to provide high spatial and temporal resolution mea-
surements. This section begins with “sensor fundamentals”, which provide the
motivation for MEMS sensors. Key choices in flow control applications are
then summarized, describing recent sensors that both highlight progress and
reveal current research topics.

II. Actuators
A. Historical Perspective
One of the most famous experiments in fluid dynamics was performed by
Schubauer and Skramstad (1948) (actually performed prior to World War II)
concerning laminar boundary layer transition on a flat plate. The authors stated
that “It was soon realized that a study of boundary-layer oscillations could be
carried out to better advantage if they were not caused by accidental disturbances
occurring in the wind tunnel but were produced by a controlled disturbance of
known amplitude and frequency at some chosen position.” After experimenting
with using sound from a loudspeaker to excite the boundary layer, they began to
develop an electrodynamic ribbon oscillator. They settled on placing a 0.002 in.
thick, 0.1 in. wide, and 3 ft long phosphor bronze strip at an approximate distance
of 0.006 in. from the plate surface. Scotch cellulose tape was used to set the stand-
off distance and hold a 12-in. segment ribbon in place in the central region of the
plate. Rubber bands were used to apply the tension required to maintain stability
of the ribbon so that it did not vibrate without excitation. The ribbon was vibrated
normal to the surface by a controllable electromagnetic force. This force was
generated by the interaction between the alternating current (ac) running through
the ribbon and the magnetic field generated by a permanent magnet on the oppo-
site side of the plate. Two hot wires were used to measure the amplification or
decay of the nominally 2-D disturbances produced by the ribbon. Interestingly,
the authors also experimented with feedback control, noting that the “system
could be made to oscillate by connecting the amplified output of the hot wire to
the ribbon.”
This was the first modern active flow control experiment that attempted to
control a small-scale flow instability (see Chapter 1). Since that time, many
researchers have developed novel actuators (and sensors) for flow control applica-
tions that leverage flow instabilities to induce large-scale changes in a flow.

B. Actuator Fundamentals
There are various ways to classify flow control actuators, including type
(e.g., fluidic, thermal, plasma, etc.) and transduction scheme (piezoelectric,
electrodynamic, electrostatic, etc.). Regardless of classification, several charac-
teristics are common to all actuators. These characteristics have analogs
with sensor design and, hence, serve as a useful basis for initial discussion.
Figure 1 defines the analogous terms and uses these to compare the character-
istics and specifications of actuators vs sensors. As shown in Fig. 1, many of
these definitions are simply reversed. For example, the input to a sensor is a
physical quantity such as shear stress or pressure, and the output is an electrical
ACTUATORS AND SENSORS 151

Actuators Sensors
• Must be intrusive • Ideally nonintrusive

• System Level Operation:


• System Level Operation:
• Input = electrical signal
• Input = flow parameter
• Output = flow disturbance
• shear stress, pressure
• Primary outputs:
• Output = electrical signal
• displacement, V, P, etc.
• primary outputs:
• Secondary outputs:
• voltage, current
• mass, mom. or vorticity flux

• Design Issues: • Design Issues:


• Application specific • Application specific

• Static Response: • Static Response:


• stroke, force, linearity, gain • max. input, dynamic range,
(output/input), etc. linearity, sensitivity (output/input)

• Dynamic Response:
• Frequency Response:
• gain (dynamic sensitivity),
• gain, phase, bandwidth, etc.
phase, bandwidth, etc.

• Sensitivity to unwanted inputs:


• May produce unwanted outputs:
• vibration, temperature, EMI
• EMI, sound, heat, etc.

• Trade-offs: • Trade-offs:
• bandwidth vs stroke • bandwidth vs gain
• stroke vs force (authority) • max input vs sensitivity

Fig. 1 Comparison between actuators and sensors. Many of the same ideas apply,
but the input and output are reversed.

quantity (e.g., voltage). For an actuator, an input electrical quantity produces


an output flow perturbation. As such, a generalized sensitivity (output/input)
can be defined for both sensors and actuators, but their dimensions and inter-
pretation are reversed.
While an ideal sensor is non-intrusive, an actuator must be intrusive by defini-
tion for it to produce a flow perturbation! An ideal sensor only responds to the
desired flow quantity, whereas an ideal actuator does not produce unwanted addi-
tional outputs (e.g., electronic or acoustic noise). The reader is encouraged to refer
back to Fig. 1 throughout the chapter.
It is generally accepted that the most important actuator design characteristics
are control authority (or stroke) and bandwidth, but these two requirements often
conflict with one another. An increase in stroke usually carries with it the penalty
of decreased bandwidth and vice versa (Mathew et al., 2006). Simply stated, the
design goal of any actuator is that it should possess adequate stroke such that it is
152 L. CATTAFESTA AND M. SHEPLAK

capable of providing a sufficiently large input to the flow system over a prescribed
range of frequencies (i.e., bandwidth) to produce the desired control effect. Quite
often the flow control objective is easily stated (e.g., attach a separated flow), but
determining a priori what constitutes a “sufficiently large input” is not. Furthermore,
the relevant physical input (e.g., velocity, mass flow, momentum, etc.) and how
the quantity should be nondimensionalized for scaling purposes is not always
clear. Part of the goal of this chapter is to address, if not answer, these questions
for several devices.
In addition to control authority, other desirable traits of an ideal actuator
include adequate bandwidth, robust operation, low cost, and energy efficiency.
An accurate model for design, scaling, and overall control-system design is
also desirable. For feedback control systems, the following three traits are also
useful: broadband spectral content, flat or non-resonant frequency response,
and linearity.
While most of the aforementioned traits are obvious, perhaps some are not. For
example, high bandwidth is often thought of as being synonymous with a broad-
band device. As shown in Fig. 2, a rotary valve actuator produces a pulsed jet with
a frequency that can be varied over a wide range—from dc to 1 kHz or more to
produce effective periodic flow excitation (Greenblatt and Wygnanski, 2000).
However, at any instant in time the frequency content consists of a single fre-
quency and perhaps its harmonics. Therefore, this actuator cannot produce a
broadband signal consisting of multiple nonharmonic frequency components, nor
does it possess the time response required to change its rotational speed to adjust
to rapid flow changes.
Whether or not these limitations are problematic depends on the application.
Figure 3 shows a block diagram of a feedback control scheme where such an
actuator cannot be used in a conventional sense (Sheplak et al., 2008). Here, the
sensor feeds back the measured dynamics of the fluid plant. In general, this signal
may consist of multiple frequency components with similar amplitudes, indicative
of a low–order but complex dynamic system. This signal is processed by the
controller, which generally demands a non-periodic waveform with multiple fre-
quency components from the actuator. It is clear that a rotary actuator, due to its
non-negligible inertia, does not have sufficient time response to produce the
desired input.

Cover
Motor Nozzles
Flexible Coupling
Bearing Bearing End Plog
Air
Inlet
Valve Body Rotor

Fig. 2 Schematic of a rotary valve actuator in which high-pressure gas is supplied to


the rotating valve. When the nozzles align between the rotating and stationary com-
ponents, a pulsed jet is supplied to the flow. The frequency of the pulsed jet depends
on the rotational speed of the motor (Greenblatt and Wygnanski, 2000).
ACTUATORS AND SENSORS 153

disturbance
noise
d n
FORWARD PATH
reference
r e x
Σ C A Σ P H Σ y
u

FEEDBACK PATH

Fig. 3 Block diagram of a feedback control scheme, in which the controller C


produces an actuator signal A to control the plant P, whose output is measured by a
sensor H. Even when there is no feedback, the actuator dynamics are often a signifi-
cant part of the overall system dynamics.

Note, however, that the actuator can still be tuned or controlled using sensor
feedback, but on a much larger (i.e., slower) time scale than that of the flow
system. But this situation implies a nonlinear dynamical process where a zero-
mean ac signal has an effect on the mean or time-averaged flowfield. Recall
that a fundamental property of a linear system is frequency preservation; a
linear system can only change the amplitude and phase of an input signal and
not its frequency.

C. Actuator Types
Our approach in this chapter is to discuss the above traits and those in Fig. 1 in
the context of several popular actuators used in the flow control community. The
operation, modeling, nondimensionalization, and scaling of many popular actua-
tors are considered. Table 1 highlights key advantages and disadvantages of each
actuator for easy reference.

1. Fluidic Actuators
a. Zero-Net Mass-Flux Actuators. Zero-net mass-flux (ZNMF) actuators,
commonly called “synthetic jets”, have been used successfully in numerous flow
control applications (Glezer and Amitay, 2002). While the device and its variants
have become extremely popular in the last decade, this type of actuator is not new.
Ingard and Labate (1950) reported steady and stroboscopic flow visualization and
quantitative measurements of essentially this device.
Figure 4 illustrates the device operation. A driver (e.g., a piston or diaphragm)
oscillates about its equilibrium position, periodically expelling and ingesting fluid
from/into the cavity through an orifice or slot. Under certain conditions, this pro-
cess can result in the formation of one or more vortex rings (Ingard and Labate,
1950; Smith and Glezer, 1998; Holman et al., 2005). These vortex ring(s) synthe-
size a time-averaged jet with a finite momentum from the surrounding fluid with
zero-net mass transfer during a cycle.
The exit plane pressure is close to ambient during expulsion but is less than
ambient during ingestion, forming the sink-like flow behavior to draw fluid back
into the actuator. This leads to a net negative mean pressure coefficient in the
exit plane of the actuator, which is problematic when considering boundary
154 L. CATTAFESTA AND M. SHEPLAK

Table 1 Summary of common flow control actuators

Type Advantages Disadvantages

Flaps – Simple design amenable to different – Product of max deflection and


frequency ranges of interest bandwidth is constant
– Can produce spanwise or stream- – Susceptible to fluid loading
wise vorticity
ZNMF – Requires no external flow source – Peak velocities typically
– Amenable to various types of limited to low to moderate
drivers and sizes subsonic speeds
Pulsed jets – Capable of high velocities with – May not be amenable to
either fast time response OR feedback control due to either
high-frequency response but frequency- or time-response
generally not both limitations
– Requires an external flow
source
VGJ – Good control authority similar to – Many adjustable parameters
mechanical vortex generators (mom. ratio, pitch and yaw
– Amenable to pulsing angles, etc.) make it non-
implementations trivial to optimize
– Requires an external flow
source
PRT – Capable of producing large – Not currently amenable to
perturbations in high-speed flows feedback control
– Requires an external flow
source
SDBD plasma – Easily installed on models (arrays) – Limited velocity output
– Can produce spanwise or stream- – Requires high voltage
wise vorticity (kilovolts)
– No moving parts
Combustion – Capable of producing large – Currently limited to relatively
perturbations in high-speed flows low frequencies – a few
hundred hertz
– Requires combustion
SparkJet – All solid-state device capable of – Currently limited to relatively
producing large perturbations in low frequencies – a few
high-speed flows hundred hertz

conditions for simulations (discussed later). In contrast, pulsed-blowing actua-


tors do not produce a mean low-pressure region because they do not have a
suction phase.
In terms of applications, a key feature of a ZNMF actuator that adds to its popu-
larity is that it requires no external flow source and, in principle, is capable of
producing complex waveforms using a variety of transduction schemes. Popular
schemes include capacitive (Coe et al., 1994), piezoelectric (James et al., 1994),
and electrodynamic (Agashe et al., 2008).
ACTUATORS AND SENSORS 155

Fig. 4 Schematic of a zero-net mass-flux or synthetic jet driven by a) piezoelectric


composite diaphragm (Glezer and Amitay, 2002) and b) electrodynamic voice-coil/
magnet assembly.

The primary advantages of piezoelectric transduction are low electrical power


requirements due to the capacitive nature of the piezoceramic, high bandwidth,
and broadband output over the dc to several kHz range (for commercially avail-
able piezoceramic disks). Their primary disadvantages include relatively low
velocity output (order of <100 m/s maximum velocity) in the standard configura-
tion shown in Fig. 4a and the requirement for resonant operation in many appli-
cations. Operating near resonance can lead to device mechanical failure, undesirable
180 deg phase shifts for feedback control (see Chapter 5), and possibly large
acoustic levels for high frequencies, resulting in a non-compact acoustic source
(see Chapter 11).
Other transduction schemes have their own advantages and disadvantages. For
example, capacitive schemes are easily amenable to standard microfabrication
techniques, although microfabricated piezoelectric and electrodynamic devices
require more sophisticated approaches (Coe et al., 1994; Madou, 1997). On the
other hand, capacitive devices are inherently nonlinear for large deflections,
whereas the others are adequately described by linear models in most cases.
Electrodynamic devices are attractive for low-frequency applications because of
their large displacement capability, but their increased weight (because of the
magnet assembly) and heat-transfer (because of heating in the resistive coil) pres-
ent formidable design challenges (Agashe et al., 2008; McCormick, 2000).
Regardless of the transduction scheme, substantial progress has been made in
the development of low-order models of sufficient fidelity for performance pre-
diction and design optimization (Gallas et al., 2003; Prasad et al., 2006). The
actuator is represented by a lumped element model (Fischer, 1955; Merhaut, 1981;
Rossi, 1988). The main assumption employed in lumped element modeling (LEM)
is that the characteristic length scales of the governing physical phenomena (e.g.,
acoustic wavelength) are much larger than the largest geometric dimension of the
device. If the lumped assumption is valid, then the temporal and spatial variations
can be decoupled. This decoupling permits the governing partial differential equa-
tions of the distributed system to be “lumped” into a set of coupled, nonlinear
156 L. CATTAFESTA AND M. SHEPLAK

ODEs. The resulting lumped parameter system is used to construct an equivalent


electrical circuit using conjugate power variables (i.e., power = effort × flow such
as voltage × current, force × velocity, and pressure × volume flow rate). The
resulting LEM provides a method to estimate the dynamic response of electrome-
chanical devices (with errors of the order of 10%) and is much simpler than a
high-fidelity, first-principles numerical simulation. The various lumped elements
in the different energy domains are represented by their respective complex-valued
impedances, Z = effort/flow.
For piezoelectric-driven ZNMF actuators, LEM is applied as follows. An ac
voltage Vac across the piezoceramic induces a mechanical strain and, hence, a
bending moment in the piezoelectric diaphragm. The diaphragm oscillates, which
produces an oscillatory output volume flow rate Qout. Since the diaphragm is used
as a piston that displaces a large volume of fluid, the frequency range of interest is
from dc to just beyond the natural frequency of the diaphragm. Higher-order
structural modes are ineffective in displacing fluid because regions with positive
displacement are offset by regions with negative displacement. The diaphragm
thus acts like an effective spring-mass-damper and is modeled by a compliance or
spring, resistance, and mass. Assuming a compressible gaseous medium, the
motion of the diaphragm either compresses/expands the fluid in the cavity or
expels/ingests fluid through the orifice or slot. The orifice is modeled by linear and
nonlinear resistors to capture viscous flow effects and noise radiation, whereas the
mass terms represent the kinetic energy of the fluid. Equations to estimate these
lumped parameters as a function of material and fluid properties and device geom-
etry are provided in Gallas et al. (2003) and the references therein.
A key conclusion from the lumped-element analysis, confirmed by experiments,
is that a ZNMF actuator is a coupled, nonlinear, two degree-of-freedom oscillator
with two characteristic frequencies. One is the natural frequency of the driver
mechanism fd (e.g., the driver or diaphragm natural frequency), and the other is the
Helmholtz frequency fH of the cavity/orifice. A common misconception is that the
two resonant frequencies of the coupled oscillator, denoted as f1 and f2, are identi-
cal to fd or fH. But this may or may not be the case and, in fact, depends on the
electromechanical coupling (Gallas et al., 2003). When the coupling is poor, then
fd is very different from fH, and two resonant frequencies indeed reduce to fd and
fH. Such is not the case when fd ~ fH. Methods to predict these natural frequencies,
f1 and f2, and the nonlinear response as a function of frequency Qout/Vac for the gen-
eral case as a function of fluid and material properties and device geometry are
presented in Gallas et al. (2003).
Others have successfully used LEM in their research. Kim et al. (2005) extended
the lumped element technique to consider the relevant case of the influence of
delays associated with long lengths of tubing, as in Fig. 5. Tang et al. (2007)
investigated various reduced-order modeling schemes and concluded that LEM
was the most accurate and suitable approach for application to full-scale aircraft
applications. Furthermore, the method is applicable to a variety of transduction
schemes. In fact, McCormick (2000) first used this technique to accurately model
voice-coil or electrodynamic ZNMF devices, followed by a more recent effort by
Agashe et al. (2008).
For more sophisticated modeling, one may turn to computational fluid
dynamics (CFD). Rumsey (2007b) summarizes the results of a workshop on
ACTUATORS AND SENSORS 157

Fig. 5 Circumferential ring of electrodynamic voice-coil actuators used in a jet


mixing experiment (Raman and Cain, 2002).

“CFD Validation of Synthetic Jets and Turbulent Separation Control”. Three test
cases of varying complexity were considered by numerous contributors: 1) a
ZNMF jet in a quiescent medium, 2) a ZNMF jet in a crossflow, and 3) ZNMF jet
control of flow over a hump model. For the isolated jet in Case 1, several issues
were noted. These included the difficulty of modeling transitional flow behavior
and 3-D effects. One simulation by Kotapati et al. (2007) showed that it is essen-
tial to match the dimensionless jet parameters using time-accurate, 3-D CFD
simulations. Despite significant simplifications in the CFD model of the actuator
geometry and diaphragm motion, good agreement between experiments and the
simulations were obtained.
The hierarchy of modeling fidelity thus varies from a complex CFD simulation
to a simple lumped element model. A CFD simulation that resolves both the
small-scale details of the actuator and the large-scale behavior of the controlled
flowfield is impractical for routine design calculations. On the other hand, a
lumped element model, while sufficient for prediction and design purposes, lacks
the fidelity required to provide an equivalent time-dependent boundary condition
for CFD that mimics the behavior of the actuator. Thus, an important current
research topic is to develop reduced-order models of a ZNMF actuator(s) suitable
for use in a controlled flowfield.
In terms of recent progress towards developing ZNMF actuators with larger
velocity output, Shaw et al. (2006) reported the development of proprietary
(General Electric) dual-bimorph piezoelectric ZNMF actuators with peak output
velocities of the order of 250 m/s. While promising, the details of the device
158 L. CATTAFESTA AND M. SHEPLAK

design are currently unavailable. Clearly, the development of high-output actua-


tors is required for application to high-speed flows. For further advancement of the
design of ZNMF devices, other relevant research issues include the development
of 1) improved nonlinear orifice impedance loss models due to oscillatory flow
through the orifice and 2) accurate grazing flow impedance models to understand
how a grazing boundary layer can load the actuator and change its output from
that measured in a quiescent medium. In the case of grazing flows, it is interesting
to note the similarities between ZNMF actuators and acoustic liners that incorpo-
rate Helmholtz resonators.
b. Piezoelectric Flaps. Another actuator that has been used successfully in a
variety of fluid dynamic applications is the piezoelectric flap actuator. As shown in
Fig. 6, these applications include control of separation (Seifert et al., 1998), turbu-
lent boundary-layer streaks (Jacobson and Reynolds, 1998; Jeon and Blackwelder,
2000), free shear flows (Cattafesta et al., 2001; Wiltse and Glezer, 1993, 1998),
and flow-induced cavity oscillations (Cattafesta et al., 1997; Mathew et al., 2006;

Fig. 6 Sample applications of piezoelectric flap actuators: active control of


a) separation, b) vortices and streaks in boundary layers, c) free shear flows, and d)
and e) for cavity oscillations. See text for references.
ACTUATORS AND SENSORS 159

Raman and Cain, 2002; Schaeffler et al., 2002). The general cantilever-beam
configuration is shown in Fig. 6e. The actuator can introduce spanwise or
streamwise vortical perturbations into the flow depending on the geometry and
orientation of the vibrating tip with respect to the local freestream flow.
Application of an ac voltage across the piezoceramic induces an asymmetric
alternating mechanical strain in the composite beam cross section due to the
piezoelectric effect. This ac strain results in oscillatory bending of the beam,
which then interacts with the flow. Composite beam modeling is treated in
Cattafesta et al. (2001). Both unimorph, with a piezoceramic on one side of the
shim, and bimorph configurations, with a piezoceramic bonded symmetrically to
both sides of the shim, are possible (Cattafesta et al., 2001; Mathew et al., 2006;
Schaeffler et al., 2002).
Typical tip displacement amplitudes range from Atip ~ O(10) to O(100) µm for
a device with a resonant frequency in excess of 1–2 kHz (Schaeffler et al., 2002)
to O(mm) when the resonant frequency is reduced to a few hundred Hz or less.
Mathew et al. (2006) discuss the necessary trade-off between the output amplitude
of the device (i.e., its “dc gain”) and its bandwidth defined by the first resonant
frequency. Analogous to an amplifier, the gain-bandwdith product is approxi-
mately constant. For example, one can achieve higher tip displacement at the cost
of a reduced usable frequency range of operation or vice versa. Mathew et al.
(2006) discuss methods to optimize the actuator geometry to meet specific
gain-bandwidth requirements. A comprehensive discussion of the performance
trade-offs inherent in the design and selection of MEMS actuators (and sensors)
such as these is presented in Bell et al. (2005).
In terms of fluid-structure coupling models, the velocity perturbation pro-
duced by the motion of the actuator depends greatly on its orientation and, in
general, requires a non-trivial fluid-structure interaction computation in which
either fluid loading or two-way coupling is considered (Kudar and Carpenter,
2007). Short of this general treatment, Cattafesta et al. (2001) discuss a simple
model for the sinusoidal free-cantilever beam case when fluid loading is not
significant and when fd/U• 1. This corresponds to the situation when the
characteristic time scale of the largest structures in the boundary layer d/U• are
much less than the period 1/f of the motion of the flap tip. When this is true, the
boundary layer is displaced by the relatively slow motion of the flap, which
enables an estimation of the local velocity perturbation induced by the flap,
u¢/U• µ Atip Red5/6/d.
c. Pulsed Jets. In this section, we highlight pulsed, as opposed to steady,
jets. It is well established that unsteady forcing has significant advantages vs
steady forcing for many flow control applications (Greenblatt and Wygnanski,
2000). Here we define a steady jet as one with negligible temporal or ac varia-
tions. A pulsed jet is an unsteady jet that is ideally either “on” or “off” and, in the
periodic case, can be characterized by a positive-definite square wave with a duty
cycle 0 £ D < 1 that indicates the percentage of the time that the jet is on. Hence,
pulsed jets have both a time-averaged or mean component and an unsteady or ac
component. We make a distinction between ZNMF and pulsed or steady jets
simply because ZNMF “jets” do not require an external fluid source. However,
most of the modeling issues are similar, with the exception of the exit pressure
boundary condition described earlier.
160 L. CATTAFESTA AND M. SHEPLAK

Fig. 7 a) Photograph of microjet array of 12 sonic nozzles, 400 mm in diameter, at


the cavity leading edge oriented normal to the incoming freestream flow. b) Schlieren
image with a horizontal knife edge with nozzle pressure ratio varying from 2.4 to 14.6
in steps of approximately 1.5 (Zhuang et al., 2006).

Pulsed jets thus require an external flow source, and the jet flow can be modu-
lated using, for example, a fast-acting solenoid valve (Bons et al., 2002), a high-
speed siren valve (Williams et al., 2006b), or a rotating orifice/slot assembly (Choi
et al., 2006). It is important to note that it is not possible to phase-lock the latter
two (rotary) types of pulsed jets with a reference signal in the flow because of
limited time response. If the time and frequency response of the solenoid valve are
sufficient, then it is possible to synchronize the valve with the reference sensor
signal. The reader should keep these ideas in mind when two implementations of
pulsed jets are described below.
d. Pulsed Microjets. The first concept is called a “pulsed microjet,” which is
a pulsed extension of steady microjets used to control separation (Kumar and
Alvi, 2006), supersonic impinging jets (Alvi et al., 2003; Lou et al., 2006),
high-speed jet noise (Arakeri et al., 2003), and flow-induced cavity oscillations
(Zhuang et al., 2006). While there is no clear definition, “micro” usually implies
sub-millimeter jet diameters, thereby significantly reducing mass flow require-
ments and enabling closely-spaced arrays. Figure 7a shows a typical installation
in the leading edge of a cavity configuration, while Fig. 7b shows schlieren images
of a microjet issuing into a quiescent medium as a function of increasing nozzle
pressure ratio. These images reveal the evolution from a subsonic jet to a perfectly
expanded jet and then to an underexpanded jet as the nozzle pressure ratio
increases. The mechanism through which the jet interacts with the flow depends,
of course, on its orientation with respect to the flow.
Recently, pulsed variants of these microjets have appeared. Ibrahim et al. (2002)
used a high-speed motor capable of achieving maximum rotational speeds of
12,000 rpm to enable unsteady forcing at 6300 Hz of compressible jets for mixing
enhancement. This frequency corresponds to a Strouhal number of 0.16, which is
close to a subharmonic of the most amplified Strouhal number. Choi et al. (2006)
recently devised a scheme to pulse the microjets using a rotating cap assembly, as
shown in Fig. 8a, via a mechanism to reduce rotational speeds. The pulsing
frequency (Hz) is

fpulsing = Nm . GR . rpm/60 (1)


ACTUATORS AND SENSORS 161

Fig. 8 Conceptual diagram of a) rotating cap actuator and b) microjet pulsing with
a phase difference (Choi et al., 2006).

where Nm is the number of microjets and GR < 1 is the gear or pulley ratio. As
shown in Fig. 8b, changing the number of slots or holes Nh in the rotating cap to
be different from Nm introduces a deterministic phase difference

ÊN ˆ
f = Á h - 1˜ ¥ 360 deg (2)
Ë Nm ¯

By changing the diameter of the holes in the rotating cap, the duty cycle can
be changed. Unfortunately, a high pulsing frequency inevitably requires large
rotational speeds, which often necessitate special bearings and cooling air to
mitigate frictional heating. They are also subject to vibration and potentially
excessive noise.
e. Vortex Generator Jets. Vortex generating jets (VJGs) have been studied
extensively in the literature (Compton and Johnston, 1992; Godard and Stanislas,
2006; Johnston, 1999; Johnston and Nishi, 1990; Khan and Johnston, 2000;
McManus et al., 1996; Selby et al., 1992; Zhang, 2003). Bons et al. (2002), for
example, have developed and applied both steady and pulsed VGJs to turboma-
chinery applications. The concept, illustrated in Fig. 9a, is to mimic conventional
vortex generators in a fluidic manner by skewing and pitching the jet axis with
respect to the freestream flow direction. As shown in Fig. 9b, this produces
co-rotating streamwise vortices that stay within the boundary layer and inhibit
162 L. CATTAFESTA AND M. SHEPLAK

a) b)
VG 2.5
U pitch 2

1.5

v-velocity (m/s)
skew 1

0.5
Surface
0

–0.5

–1
–5 –2 1 4 7 10 13
z/D

Fig. 9 a) Schematic of a vortex generator jet, and b) wall normal velocity measure-
ments superimposed on streamwise velocity contours showing streamwise vortices for
ujet /U• = 4 (Bons et al., 2002).

separation via momentum exchange inside the boundary layer. The authors have
shown that, by using a fast-response solenoid valve with a response time of
approximately 1 ms and a bandwidth of 250 Hz, they can significantly reduce the
duty cycle, which reduces the required mass flow rates, without jeopardizing
performance benefits. However, the resulting unsteady, 3-D turbulent flowfield
requires time-accurate, high-fidelity CFD simulations to accompany experiments
in order to truly understand key mechanisms and optimize adjustable parameters
for effective control.
f. Hartmann and Powered Resonance Tubes. The above pulsed jet actuators
ultimately fall short of being able to force high-frequency instability mechanisms
due to either the finite response time of the valves or physical rotational limita-
tions. A different type of fluidic actuator with high-frequency forcing capability is
the power resonance tube (PRT) or whistle which takes advantage of acoustic
resonance. These actuators are an adaptation of the Hartmann tube first discovered
by Hartmann and Trolle (1927) and have been subsequently adapted by several
researchers for active flow control applications (Kastner and Samimy, 2002;
Raman et al., 2001, 2004a, b; Sarpotdar et al., 2005; Stanek et al., 2002a, b).
The basic concept, illustrated in Fig. 10a, is to align a jet issuing from a
converging nozzle with the axis of a closed tube of length L. An array of PRTs
is shown in Fig. 10b for control of flow-induced cavity oscillations. A cycle is
established in which incident compression waves generated by the impinging
jet travel down the tube and reflect from its end as compression waves. These
waves then reflect as expansion waves at the tube entrance, which subsequently
travel down the tube length and reflect from the end as expansion waves.
When these reach the entrance to the tube, they reflect as compression waves,
and the cycle begins anew. Raman et al. (2001) describe how a simple quarter-
wave resonator model describes the behavior for long tubes, fres = c0/4L.
Additional progress has been made in the development (Raman et al., 2004b;
ACTUATORS AND SENSORS 163

Fig. 10 a) Schematic of a Hartmann tube, showing a converging nozzle issuing a jet


of diameter d into a tube of length L, separated by a distance Dx; and b) a bank of
powered resonance tubes mounted at the leading edge of a cavity (Stanek et al.,
2002b).

Sarpotdar et al., 2005), higher-order modeling (Kerschen et al., 2004), and


numerical simulations of these high-bandwidth actuators (Cain et al., 2004),
including extensions to lower frequencies by incorporating a Helmholtz resona-
tor to reduce the required tube length.
The main attraction of these actuators is their ability to generate large high-
frequency excitations suitable for control of high-speed flows—hence the name
“HIFEX.” While theories have been proposed about the mechanism responsible
for control (Stanek et al., 2002a, b, 2003), it is important to emphasize that these
devices impart a non-zero mass and momentum to the flow, O(1%) of the
freestream mass flow rate. These disturbances affect not only the targeted high-
frequency instabilities but also the mean flow (Ukeiley et al., 2004). Because the
disturbances are large, this has two possible effects. First, the mean flow instabil-
ity characteristics are modified. Second, the large disturbances may produce
nonlinear interactions between multiple frequency scales.
A second point worth noting is that these actuators possess high bandwidth and
are tunable (by adjusting, for example, the tube length) but do not possess fast
time response. Hence, the excitation signal frequency content cannot be easily
controlled in a rapid fashion suitable for dynamic feedback flow control. For this
to be possible, the time response of the actuator would have to be significantly
smaller than the times scales of the fluid dynamic processes. Furthermore, as with
rotary-valve based devices, these actuators cannot be phase-locked to a measured
signal in the flow.
g. Scaling of Fluidic Actuators. It is important to understand how these
fluidic actuators are characterized and scaled. Two dimensionless parameters are
commonly used in the literature for this purpose: unsteady blowing

Ê rjet ujet ˆ Ajet m jet


Bc = Á ˜ = (3)
r U
Ë • • ¯ ref A r • • Aref
U

and momentum coefficients

/
cm = rjetU 2jetrmsAslot (q• Aref) (4)
164 L. CATTAFESTA AND M. SHEPLAK

where q• = 0.5r•U•2 and Aref is an appropriate reference area (e.g., cavity open
area = length × width or airfoil planform area = chord × span). If the devices are
not zero-net mass flux, there will be a mean component in addition to the unsteady
value.
Note that the literature is full of variations on these parameter definitions, which
emphasizes the need for actuator designers to specify their terminology clearly.
For example, in Bc, ujet may represent an amplitude or an rms quantity. Further-
more, in both Bc and cm, the velocity may represent a spatial average across the slot
or just a single measurement at the slot center. It is also important to note that
these actuators are often characterized in a quiescent medium and not in situ.
Because the grazing flow imposes an additional impedance on the device, the
output of the actuator during application may be substantially different from the
“benchtop” calibration (Utturkar et al., 2002). Hence, benchtop calibrations can
only be regarded as estimates of the true actuator output.
In addition to the above quantities, the Strouhal number of the targeted fluid
dynamic instabilities, Sr = fLref /U•, and the relative size of the actuator compared
to some flow scale (e.g., hslot /d ) are often very relevant. The benefit of using
dimensionless parameters in actuator design is that the parameters provide a
means to evaluate the suitability of an actuator as a function of scale (i.e., as the
application migrates from a small-scale wind tunnel to a full-scale test).
As an example, consider the control of flow-induced cavity oscillations. If a
1/20th scale model is tested and U• is approximately the same in the scaled and
full-scale cases (e.g., for Mach number similarity), then the Strouhal number scal-
ing dictates that the characteristic frequency of the cavity tones will reduce by a
factor of 20 in the full-scale application. This has implications concerning the
required actuator bandwidth. Perhaps a different actuator is more suitable for the
full-scale test?
Furthermore, assuming 2-D spanwise slots are used, given that U jetrms =
÷ (cm q• Lref)/(rjet hslot), we must decide whether we wish hslot /d or hslot /Lref to remain
constant. If the latter is chosen and density is constant, then ujetrms/U• is unchanged.
But in this case, hslot /d may not be the same at full scale as in the wind tunnel tests
due to incomplete (Reynolds number) similarity, which may change how the actua-
tor couples with the flow instability. This exercise does not provide a solution but
merely points out some issues that must be considered.

2. Non-Thermal Plasma Actuators


The potential advantages of plasma actuators for flow control are widely recog-
nized in the literature. Here, we restrict our attention to the most popular type, the
so-called non-thermal plasma actuators, because of their ability to modify the flow
near atmospheric conditions over a control surface. These actuators have no moving
parts, thus providing rapid time response. Moreau (2007) and Corke et al. (2007)
provide excellent reviews of these devices. A popular variant of these devices
discussed below is the single dielectric barrier discharge (SDBD) actuator.
As shown in Fig. 11a, a SDBD actuator consists of one dielectric coated elec-
trode exposed to the flow surface and a second electrode embedded in an insulat-
ing layer a short streamwise distance from it (Roy, 2005). A high-voltage ac input,
O(kV), is supplied to the electrodes, which results in an asymmetric electric field.
ACTUATORS AND SENSORS 165

a) b)
Electrohydrodynamic (EHD) 0 Mean 1
Flow Body Force
Actuation (4.7, 22.5)
Fluid ef
Using
EHD Force
Electric Field Lines
frf

(3, 0.2)
No
Rotating Dielectric
Component Surface
Discharge 0 Fluctuating RMS 1
Dielectric
f=0 (4.7, 22.5)
ed

(3, 0.2)

Fig. 11 a) Schematic of a single SDBD actuator (Roy, 2005), and b) PIV velocity
measurements (units of m/s) of a single SDBD actuator with Efield = 8.47 kVp-p at 3 kHz
(Wilkinson, 2003).

Above the breakdown voltage VB, a visible plasma is produced as the air ionizes.
According to Paschen’s (1889) law, the breakdown voltage for a particular gas
depends on the product of the gas pressure P and the distance d between the
electrodes. For any gas there is unique value of this P · d product where volumetric
ionization is the maximum (e.g., for air P · d = 5.7 Torr-mm).
Unfortunately, for flow control applications near atmospheric pressure, the
allowable electrode spacing necessary for maximum volumetric ionization is
d = 7.5 µm. In most applications, specifically in high-speed air vehicles, this is an
impractical limitation. A solution to this limitation comes from the recent devel-
opment of RF glow discharge using an ac voltage potential across the electrodes
(Kanda et al., 1991). A homogeneous glow can be maintained at 3–20 kHz RF
with rms voltages between 2 and 15 kV. A critical criterion for such discharge in
air is to meet the electric field requirement of ~30 kV/cm.
The electric field gradient produces an electrohydrodynamic body force BE on
the bulk air, resulting in a weakly ionized plasma (Roth et al., 2000). This body
force is analogous to a pressure gradient for a 1-D case

1 d Ê1 ˆ
BE = 2
e —◊ Efield æ1æ
D
Æ e E2
2 0 dx ÁË 2 0 field ˜¯ (5)

units of pressure

The actuator body force can be calculated offline for a given configuration and
then incorporated into the momentum equation for a numerical simulation of flow
control (Corke et al., 2007).
In a quiescent medium, the net effect is to induce a tangential wall jet, as shown
in Fig. 11b from Wilkinson (2003). Several papers highlight the theoretical and
numerical modeling efforts of this barrier discharge. These efforts span a range of
phenomenological to first-principle-based methods (Enloe et al., 2003; Massines
et al., 1998; Roy and Gaitonde, 2005; Roy et al., 2006; Shyy et al., 2002).
Key research issues with these devices include the control authority or lack thereof
as the flow velocity increases, their suitability in high relative humidity environ-
ments, and a fundamental understanding and characterization of plasma-induced
166 L. CATTAFESTA AND M. SHEPLAK

flow control in real gas chemistry. Reported experimental data is mainly focused on
measurements of velocity distribution about the SDBD actuator and spatially aver-
aged thrust. There are currently no experimental data available that provide a detailed
understanding of ionized air chemistry, the distribution of plasma density, or the
spatially and temporally resolved force density. These issues must be addressed in
order to transition these actuators from the laboratory to flight.

3. Combustion Actuators
This section describes novel, high-output, small-scale combustion-based fluidic
actuators for flow control applications (Crittenden et al., 2001, 2004; Cutler et al.,
2005). The combustion-driven actuator jet is a pulsed jet that is produced by the
ignition of a mixture of gaseous fuel and oxidizer in a small (cm3 scale) combus-
tion chamber, as shown in Fig. 12a. The cycle begins with the injection of pre-
mixed fuel and oxidizer into the combustion chamber, displacing the remaining
combustion products from the previous cycle. An integrated small-scale spark
ignites the mixture, and a combustion process ensues that typically lasts several
milliseconds (depending on the type of fuel, mixture ratio, and physical sizes of
the combustor and orifice). Combustion results in a rapid pressure rise in the
chamber and the ejection of a pulsed high-speed jet through one or more orifices.
The operating frequency can be varied by controlling the flow rate of the fuel/
oxidizer and the ignition frequency. Frequencies greater than 150 Hz have been
achieved with chamber pressures of up to 5 atm, and these devices are capable of
producing sonic velocities at the jet orifice. Significant jet penetration into a cross-
flow at Mach numbers up to 0.7 has been demonstrated (Crittenden et al., 2001).
Examples of applications include transient separation control and counterflow
configurations, which are attractive for controlled efficient mixing.
The primary attraction of this device is its high-velocity output. However, due
to the finite time duration associated with the combustion cycle, the device is

a) b) 3

High Velocity
Pulsed Jet

Combustion
Chamber 2
Pc/Patm

Periodic
Spark Ignition

Flow
Regulator
Premixed
Reactants 1

0 1 2 3 5
Time (ms)

Fig. 12 a) Schematic illustration of combustion-driven jet actuator, and b) phase-


locked schlieren images and pressure trace for 1 cm3 combustion chamber with stoi-
chiometric hydrogen mixture and 1.27 mm orifice diameter (Crittenden et al., 2001).
ACTUATORS AND SENSORS 167

limited in its current form to frequencies less than approximately 1 kHz. In addi-
tion, although similar in many respects to a synthetic jet, this device is not a ZNMF
device due to the requirement for small but non-zero reactant flow. Hence, imple-
mentation is more complex. Finally, it should be noted that the device is not easily
amenable to feedback control applications because of its discrete pulse behavior.

4. SparkJet Actuators
A solid-state variant of the combustion actuator that does not require fuel or
valves is the SparkJet actuator being developed at the Johns Hopkins Applied
Physics Laboratories (Cybyk et al., 2003, 2004, 2005; Grossman et al., 2003, 2004).
A schematic of this device is shown in Fig. 13a. The body is made from an electrical
insulator such as a ceramic, and the actuator consists of three interchangeable com-
ponents: an anode, a sharp cathode, and a grid. The device is compact, with typical
cavity volumes from 28 to 52 mm3 and orifice diameters from 0.08 to 0.5 mm.
The principle of operation is straight forward. The discharged current flows
from the cathode to the anode and is initiated by a small cathode-to-grid discharge.
A large chamber pressure is thus generated via rapid gas heating inside the
SparkJet. The high-pressure gas exhausts through the chamber orifice at high velo-
city. Although the SparkJet device is configured similarly to a vacuum tube triode,
the operation of the device is significantly different. The cathode is not initially
heated and breakdown is induced by high electric fields. Also, the discharge occurs
at ambient conditions instead of in a vacuum tube.
A single cycle of SparkJet operation consists of three distinct stages: energy
deposition, discharge, and recovery, as shown in Fig. 13b. In the energy deposition
stage, a capacitive element connected in parallel between the anode and cathode
is charged with increasing voltage until breakdown occurs, and the electric arc
discharge rapidly heats and compresses the bulk chamber gas. As the capacitor

Fig. 13 a) Schematic illustration of solid-state spark jet actuator and b) schematic


illustration of three distinct stages of a cycle (Cybyk et al., 2003, 2004, 2005; Grossman
et al., 2003, 2004).
168 L. CATTAFESTA AND M. SHEPLAK

discharges, the voltage ultimately drops below that required to sustain the
discharge, and the arc quenches. During the discharge stage, the gas is expelled
through a choked small orifice at high speeds until the orifice eventually unchokes,
and the exhaust velocity decreases to zero. During the recovery stage, the chamber
cools and draws fresh air from ambient into the chamber.
Computational (e.g., Cybyk et al., 2003) and various experimental techniques
have been employed to model and investigate the performance of the SparkJet
actuator for control of supersonic flows over external surfaces. A detailed under-
standing of the dependence of the jet velocity, temperature, and time constant as
a function of, for example, energy deposition, cavity volume, and orifice dimen-
sions is required for intelligent design. Typical values of these parameters are an
output velocity of 450 m/s, a jet temperature of 500 K, and cycle times of 1–2 ms
for an energy deposition of 20 mJ. The main disadvantage of the device in its
current form is the relatively long cycle times, which limit the ability of the device
to specifically target high-frequency instabilities in high-speed shear flows.

III. Sensors
A. Sensor Basics
For realistic 3-D flow-control applications, measurements of the wall-stress
components are more practical than flowfield measurements of velocity or thermo-
dynamic quantities. The normal stress for flows obeying Stokes’ hypothesis is equal
to the scalar thermodynamic pressure. The tangential vector components are decom-
posed into streamwise and spanwise viscous wall shear stresses. For the applications
of pressure-drag and skin-friction drag reduction, a measurement of these quantities
has the added benefit of being directly related to the control objective. The question
arises about the advantages of pressure vs shear stress sensing for feedback flow
control. As mentioned above, wall shear stress tw is a vector property that offers the
advantage of flow direction over scalar pressure. Recent studies indicate that,
depending on the application, it may be advantageous to use skin friction alone or in
combination with pressure for state estimation (Aamo et al., 2003; Alam et al., 2006;
Rathnasingham and Breuer, 2003; Surana et al., 2006).
Some important specifications for pressure and wall shear stress sensors are the
dynamic range, bandwidth, and sensitivity (see Fig. 1). For the purposes of illus-
tration each of these characteristics is defined for a pressure sensor, but the defini-
tions hold for shear stress sensors as well. The dynamic range in units of decibels
is defined as 20 log(pmax/pmin). The upper limit of the dynamic range, pmax, is the
maximum rms pressure that can be linearly sensed with a harmonic distortion
level below a specified value (typically 3% or 10% for microphones). This limit is
usually determined by transduction and/or interface electronics nonlinearities.
The lower limit of the dynamic range, pmin, is the minimum detectable rms pres-
sure determined by the noise floor of the sensing system (Brüel and Kjær, 1996).
The bandwidth is defined as fmax – fmin, where fmax and fmin are the maximum and
minimum frequencies, respectively, at which there is a ±3 dB deviation from the
flat band magnitude of the transducer frequency response function.
All sensors possess finite inertia (mechanical or thermal) that inherently limits
their dynamic response. To first order, all mechanical sensors can be approximated
ACTUATORS AND SENSORS 169

as a second-order system up to just beyond the first____mechanical resonance, so fmax


scales with the first resonant frequency, fmax @ ÷k/m , where k is the effective
stiffness and m is the effective mass of the system (Merhaut, 1981). For sensors
that measure a mean flow quantity (i.e., a pressure sensor), fmin = 0 Hz. For ac
sensors such as microphones, fmin is determined by the cavity/vent acoustics and
interface electronics (Brüel and Kjær, 1996). As shown in Fig. 1, the sensitivity of
a sensor is defined as the output (usually voltage) divided by the physical input
(pressure or shear stress). For mechanical sensors, sensitivity is equal to the com-
pliance of the system or the inverse of the stiffness, S = 1/k. As was discussed
earlier for actuators, there is a trade-off between the sensitivity and bandwidth of
_____
a mechanical sensor due to the sensitivity–bandwidth product, Sfmax  ÷1/km ,
which is approximately constant for a given transducer architecture. Micro-
machining technology enables the development of thin, low-mass, compliant
MEMS sensors possessing higher sensitivity-bandwidth products relative to con-
ventional sensors (Naughton and Sheplak, 2002).

B. Specific Requirements of Closed-Loop Flow Control Strategies


Whether used for feedback control or state estimation, an ideal sensor and
associated interface electronics (i.e., the sensing system) possess several desirable
traits (Sheplak et al., 2008). These are described in the following sections.

1. Spatial and Temporal Resolution


The sensor system should possess sufficient spatial and temporal resolution to
accurately capture the relevant flow physics to be controlled. For example, in the
case of pressure fluctuations in a turbulent boundary layer, insufficient spatial
resolution results in spatial averaging and an attenuation of the turbulent spectra
(Corcos, 1963). Inadequate temporal resolution leads to a low-pass filtering of the
measured data. The most stringent resolution requirements are for the case of
turbulent boundary-layer control, where it may be desirable to sense at the
Kolmogorov scale possessing spatial length scales of interest O(100 mm) or less
and the required bandwidth can be O(1 kHz) or more (Naughton and Sheplak,
2002). Fortunately, many flow control applications, such as separation control,
are dominated by large-scale structures and do not require accurate sensing with
resolution of the Kolmogorov scales.

2. Non-Intrusive Nature
Ideal sensors must be non-intrusive and should not actuate the flow. Pressure
and wall shear stress sensors must be flush-mounted to the aerodynamic surface.
The degree of roughness of the sensor package that is considered non-intrusive is
a function of the flow. For a turbulent boundary layer, the sensor package should
be hydraulically smooth; thus the roughness is confined to the viscous sublayer.
_____
For a roughness height hk, this requires that u*hk /n < 5 where u* = ÷tw /r• is the
friction velocity (Winter, 1977). This requirement may be even more stringent
for transition experiments and relaxed for separated flows. If a sensor is mechani-
cally non-intrusive, it still may transfer energy to the flow, thus behaving like an
170 L. CATTAFESTA AND M. SHEPLAK

actuator. For example, it is well known that thermal-based wall shear stress sen-
sors locally heat the flow, thus perturbing the velocity profile and altering skin
friction (Appukuttan et al., 2003).

3. Control System Issues


As shown in Fig. 3, the sensor feeds back the measured dynamics of the system.
Therefore the sensor dynamics should not alter the true dynamics of the system.
The ideal sensor system should possess known constant gain and phase responses,
because large phase lags can lead to controller instabilities (Ogata, 2001). The
sensor should also respond in a linear manner to the largest expected fluctuations
to avoid controller instabilities. This is especially important for separation control
applications where large fluctuation levels may be present, which can destabilize
the controller (Banaszuk et al., 2006a; Ogata, 2001; Tian et al., 2006a, b). Further-
more, the sensor should be immune to unwanted inputs such as electromagnetic
interference, vibration, etc. One example in which sensitivity to unwanted inputs
can adversely affect a control system is the direct feedthrough of an actuator signal
to a sensor (e.g., via acoustic vs hydrodynamic path of a ZNMF actuator). This
direct feedthrough can manifest itself as a notch in an experimentally obtained
transfer function and can lead to undesirable pole-zero cancellation and poor
controller performance (Banaszuk et al., 2006a).

4. Practical Implementation Issues


There are several requirements for a physically realizable and economically fea-
sible sensing system. In many applications, flow control inherently involves sensing
and actuating a distributed field. Therefore the sensing system may require array
capabilities. For such applications, gain and phase matching between the sensors is
important to avoid sensing errors. Furthermore, the electronic powering and signal
read-out requirements for sensor arrays must be considered when choosing a trans-
ducer. From a practical perspective, the sensing system should be sufficiently robust,
consume minimal power, and be economically feasible from a cost perspective
(Sheplak et al., 2008). The inherent small physical size, the ability to be placed in
close proximity, and batch-fabrication characteristics of microfabricated transduc-
ers offer the potential to meet the demands of flow control applications.

C. Microelectromechanical Systems
MEMS technology extends silicon-based integrated circuit manufacturing
approaches to enable miniature engineering systems (Madou, 1997). This enabling
technology provides the opportunity to create electromechanical sensors and actua-
tors possessing performance that is often not possible with conventional macroscale
manufacturing techniques. MEMS technology has many existing applications in
industry and has spawned a growing fluid mechanics research community (Ho
and Tai, 1998). From the perspective of fluid mechanics measurement instrumen-
tation, the small physical size and reduced inertia of microsensors vastly improve
both the temporal and spatial measurement resolution relative to conventional
sensors. The first silicon sensors began to appear shortly after the discovery of the
piezoresistance effect by Smith (1954) at Bell Laboratories. Realizing the potential
ACTUATORS AND SENSORS 171

performance advantages of microfabricated sensors via MEMS technology, a


number of companies, such as Kulite (http://www.kulite.com) and Endevco
(http://www.endevco.com), developed commercial dynamic pressure sensors that
are widely used in the fluid dynamics community. The development of a com-
mercially available MEMS shear stress sensor has been more elusive. A number
of researchers have presented an assortment of micromachined pressure and shear
stress sensors for fundamental fluid mechanics measurements and flow-control
applications that are at various stages of technical maturity (Kälvesten et al., 1994;
Löfdahl and Gad-el-Hak, 1999). Table 2 summarizes the advantages and disad-
vantages of the various sensors discussed below.

Table 2 Summary of common flow control MEMS sensors

Type Advantages Disadvantages

Piezoresistive – Commercially available – Provides scalar information


dynamic pressure – Reasonably robust – Low sensitivity
sensor or – Excellent spatial resolution – High noise floor
microphone – Large bandwidth (>100 kHz)
Capacitive – Commercially available – Provides scalar information
microphone – High sensitivity – Designed for audio applications,
– Excellent spatial resolution so upper end of dynamic range
– Low noise floor may be insufficient
Thermal shear – Reasonable bandwidth – Indirect measurement, qualita-
stress sensor – Time-resolved information tive in nature
– No directionality
– Thermal perturbation to flow
Microfences – Reasonable bandwidth – Indirect measurement, must be
– Time-resolved information in sublayer
– Poor spatial resolution
Micropillars – Arrays – Indirect measurement, must be
– Excellent spatial resolution in sublayer
– Bandwidth and time-resolved
data may be a challenge for
certain flows
– Optical access required
– Flow perturbation?
Laser-based – Excellent spatial resolution – Indirect measurement, must be
velocity gradient in sublayer
measurement – Bandwidth and time-resolved
data may be a challenge for
certain flows
Floating element – Direct measurement of wall – Currently a research area, not
sensors shear stress reduced to practice
– Time resolved
– Excellent spatial and
temporal resolution
– Array capability
172 L. CATTAFESTA AND M. SHEPLAK

Fig. 14 Cross-sectional schematic of a generic microphone structure.

1. Unsteady Pressure Sensors


Microphones and unsteady pressure sensors are electro-mechanical-acoustic
transducers that transform or modulate acoustic energy into electrical energy
(Fig. 14). The main difference between the two is that a microphone is an ac
measurement device that does not respond to changes in the mean pressure due to
its vent structure, whereas unsteady pressure transducers are dc measurement
devices that respond to both static and dynamic pressure changes. Typically, pres-
sure sensors possess lower sensitivities and higher noise floors than microphones
because they must be able to measure large absolute pressures linearly. Micro-
phones, however, only measure pressure fluctuations which are usually much
lower than absolute pressures. As a result, microphone diaphragms are more
compliant, leading to a higher sensitivity. For a given dynamic range, the noise
floor is therefore reduced compared to an absolute pressure transducer.
Although many different transduction principles have been employed, all of
these sensors are based on the measurement of a pressure-induced structural deflec-
tion. MEMS microphones and dynamic pressure sensors are particularly valuable
to the fluid mechanics community because they may reduce flow disturbances
relative to conventional sensors and can achieve fine spatial resolution without using
a pinhole mounting scheme. Most existing silicon microphones were designed for
audio applications such as hearing aids and lack the dynamic range and bandwidth
requirements for fluid mechanics applications (Scheeper et al., 1994). However,
turbulence measurements and flow control applications require measurement
capabilities beyond the threshold of pain (~120 dB ref. 20 mPa = 0.0029 psi),
which is the upper end of the dynamic range or distortion limit for audio applica-
tions (Löfdahl and Gad-el-Hak, 1999).
The most popular MEMS microphone type employs the capacitive-detection
scheme and typically requires on-chip electronics to minimize the effects of para-
sitic capacitance due to the high impedance O(1 GW) of the transducer (Scheeper
et al., 1994). Existing piezoelectric schemes are limited by insufficient dynamic
range due to high noise floors (Scheeper et al., 1994). The piezoresistive transduc-
tion scheme which consists of measuring the strain on or near the surface of a
deflected diaphragm is less expensive to develop, simpler to fabricate, and more
robust than a capacitive device. In addition, the relatively low impedance O(1 kW)
ACTUATORS AND SENSORS 173

of piezoresistive sensors enables off-chip signal measurement without significant


loss in performance. These benefits are reflected in the fact that silicon micro-
machined piezoresistive microphones are the oldest silicon microphones (Burns,
1957) and have been commercially available for decades. Furthermore, only
piezoresistive MEMS transducers have been applied as fluid mechanics and
aeroacoustics research tools (Arnold et al., 2003; Huang et al., 2002; Kälvesten
et al., 1994; Löfdahl and Gad-el-Hak, 1999; Scheeper et al., 1994).

2. Wall Shear Stress


Conventional measurement technologies are not capable of obtaining accurate
fluctuating 3-D wall shear-stress data required for both fundamental fluid mea-
surements and some flow control applications (Fernholtz et al., 1996; Haritonidis,
1989; Löfdahl and Gad-el-Hak, 1999; Naughton and Sheplak, 2002; Winter,
1977). For example, conventional floating element sensors typically possess
spatial resolutions of O(1 cm) and resonant frequencies O(100 Hz). This situation
has motivated the development of micromachined sensors to overcome some of
the traditional limiting factors associated with conventional techniques.
MEMS shear stress sensors, like their conventional counterparts, are broadly
classified as direct and indirect sensing techniques (Fernholtz et al., 1996;
Haritonidis, 1989; Löfdahl and Gad-el-Hak, 1999; Naughton and Sheplak, 2002;
Winter, 1977). The former directly measure the shear stress acting on the sensor
surface. This is typically achieved by employing a “floating element” balance.
Indirect techniques require an empirical or theoretical correlation, typically valid
for very specific conditions, to relate the measured property to the wall shear stress.
The MEMS community has produced a variety of different indirect transduction
schemes, such as hot-film sensors, micro-optical systems to measure near-wall
velocity gradients, mechanical microfences, and micropillar devices. The respec-
tive advantages and disadvantages of these devices for flow control applications are
summarized below. The interested reader can find the detailed reviews and asso-
ciated references for all sensors discussed in several MEMS shear stress sensor
review papers (Naughton and Sheplak, 2002; Sheplak et al., 2004).
a. Indirect MEMS Sensors. All indirect shear stress sensors require a corre-
lation between a measured flow property (heat transfer, velocity profile in the
sublayer, etc.) and the desired wall shear stress. Typically, the calibrations for
these devices are only valid under very specific flow conditions. For example,
laser-based optical MEMS (MOEMS) sensors measure the velocity gradient in
the viscous sublayer of a boundary layer and relate that to the wall shear stress
(Fourguette et al., 2001). Microfence sensors infer the wall shear stress by placing
a small fence within the viscous sublayer of a turbulent boundary layer (Schöber
et al., 2004). The static pressure drop across the upstream and downstream side of
the fence is then related to the wall shear stress via a calibration curve for a known
velocity profile. The extension of these techniques to realistic complex 3-D flows
is an unresolved challenge. Specifically, both of these techniques implicitly
assume that the viscous sublayer is a universal feature of the flow to be measured
and the measurement volume resides within this layer. More recently, micropillar
sensor arrays have demonstrated promising results for measuring wall shear stress
distributions (Brücker et al., 2005). Micropillars consist of miniature cylinders
174 L. CATTAFESTA AND M. SHEPLAK

with diameters of a few microns and lengths of a few hundred microns. The
micro-pillars are attached to the surface and extend into, and sometimes beyond,
the viscous sublayer. The crossflow results in a nonuniform loading that causes
deflection of the pillar which in turn is measured via an optical technique. The
absolute calibration of these devices in complex 3-D flows remains unresolved.
Thermal sensors are temperature-resistive transducers (see, for example, Liu
et al., 1999). The sensing element is heated to a temperature greater than the fluid
temperature which generates a thermal boundary layer d T (x) within the velocity
boundary layer d(x) (Fig. 15). As the temperature of the sensor varies with con-
vective heat transfer changes in the flow environment, so does the resistance and,
hence, the Joulean heating rate. This heating rate is then related to wall shear
stress. Thermal-based shear stress sensors possess several additional limitat-
ions when used for quantitative wall shear stress measurements. Specifically, the
uncertainty of the dynamic response of these thermal techniques—due to heat
conduction to the wall, calibration difficulties, flow perturbation due to heating,
and errors in response to large fluctuations with respect to the mean (~40%)—
have not been quantified (Naughton and Sheplak, 2002). There is considerable
evidence that the uncertainty of thermal sensors can be quite large in gas flow
applications. In particular, a recent computational study suggests that perturba-
tions due to heat transfer from the sensor to the flow can alone result in mean shear
stress errors of 5% or greater (Appukuttan et al., 2003). In addition, a single ther-
mal sensor is unable to discern the direction of the wall shear stress, thus limiting
its usefulness in the vicinity of separating and reattaching flows. Conversely,
thermal sensors can be useful to infer the locations of transition, separation, and
reattachment (Bertlerud, 1998). Their value as a qualitative measurement tool for
feedback control remains an open question. However, their quantitative use is
inadvisable due to the large uncertainties summarized above.
b. Direct MEMS Sensors. Direct sensors measure the integrated force
produced by the wall shear stress on a flush-mounted, movable, “floating”
element (see, for example, Schmidt et al., 1988). The floating element is either
attached to a displacement transducer or is part of a feedback force-rebalance

d (x)

Thin film sensor

dT (x)

Chip

Fig. 15 Side-view schematic of an indirect thermal wall shear stress sensor of size L
illustrating the viscous d(x) and thermal d T (x) boundary layers.
ACTUATORS AND SENSORS 175

Fig. 16 Plan-view and side-view schematics of a floating element wall shear stress
sensor illustrating the sensor dimension, gap height g, and effective spring constant k
provided by the beam-like tethers of the device.

configuration (Fig. 16). Floating element techniques appear to be better suited


for obtaining quantitative, time-resolved data, provided that a stable, low-noise
transduction scheme can be developed that is immune to both EMI and trans-
verse motions. From a packaging perspective, the sensor system must possess
backside electrical or optical interconnects to provide a truly flush-mounted
device. The robustness of the sensors to debris must also be addressed by cover-
ing the sensor gaps. Finally, the transduction scheme should permit the ability
to realize arrays O(10s) to O(100s) of sensors to map wall shear stress fields.
For array applications, the issues of sensor/electronics powering and signal
read-out are of critical importance.

IV. Concluding Remarks


This chapter has reviewed popular actuator and sensor technologies for flow
control applications. In the case of actuators, one may summarize the current state
of affairs by stating that no one actuator technology stands out as the best for all
applications. Some, such as ZNMF fluidic actuators, are well suited for closed-
loop applications because of their broadband output and fast time response.
However, to date, these inevitably suffer either from insufficient control authority
or bandwidth limitations in high-speed flow applications. Others, such as PRTs,
are capable of producing high-frequency perturbations of sufficient amplitude for
high-speed flow applications, but lack the controllable broadband output and/or
fast time response required for conventional feedback flow control.
176 L. CATTAFESTA AND M. SHEPLAK

It is clear that the community will and should continue research on the develop-
ment of high-performance actuators. For progress to continue, renewed emphasis
should be placed on modeling such devices to understand their capabilities, limi-
tations, and their cost. Indeed, a consideration of the cost of active control should
be considered for each application, and this cost metric could be related to the
actuator power consumption or some measure of the control input magnitude
(e.g., momentum coefficient). An intelligent actuator choice cannot be made
without open discussions of these relevant issues.
With regard to sensors, attention was limited to wall-bounded pressure and shear
stress sensors because of their practical relevance in feedback control. The push
for high-performance, large bandwidth, and nonintrusive devices points squarely
towards MEMS-based sensors and sensor arrays. MEMS pressure sensors have
made great strides in this regard. However, all existing MEMS shear stress sensors
are still fairly immature and require further development to become reliable mea-
surement tools for feedback control. In particular, the question of “what is good
enough?” for effective feedback control remains an open question that must be
addressed for each application. Finally, as noted in Sheplak et al. (2008), while
microfabrication technology is fairly well-established, flow sensor research and
development is currently insignificant when compared to actuator-related research
and theoretical, computational, and experimental flow control efforts. Furthermore,
sensor research is usually treated as a separate problem. We believe that continued
progress requires that all aspects of a flow control system, including sensors, actua-
tors, fluid dynamics, and controls, must be treated together.
Chapter 7

Modeling and Simulation

Christopher L. Rumsey* and R. Charles Swanson*


NASA Langley Research Center, Hampton, Virginia

I. Introduction
Active flow control has rapidly become a major topic—almost a discipline in
and of itself—within the field of fluid dynamics. As such, a large amount of
experimental and computational work has occurred, especially within the last 20
years. The reader is referred to several review articles on the topic (Bewley, 2001;
Collis et al., 2004; Greenblatt and Wygnanski, 2000; Stanewsky, 2001). The driv-
ing force behind the interest in active flow control is the fact that many companies
and research organizations see great potential gains from its use for a variety of
different aeronautical and naval applications.
Most of the computational work has been done within the broad category of the
Navier–Stokes equations for Newtonian continuum fluid motion. Many computa-
tional methods for solving various forms of these equations have been developed
since the advent of computational fluid dynamics (CFD), and these have been
readily applied to flow control applications by the CFD community with varying
levels of success. The purpose of this chapter is to summarize many of the meth-
ods used for modeling and simulation of flow control applications. In particular,
we attempt to strike a balance between a summary of the equations, a review of
some of the numerical techniques used, and a discussion of issues that have arisen
in CFD validations.
This chapter is organized as follows. In Sec. II various CFD methodologies are
described. These include direct numerical simulation (DNS) in Sec. II.A, large-
eddy simulation (LES) in Sec. II.B, and Reynolds-averaged Navier–Stokes
(RANS) in Sec. II.C. Each of these sections describes the relevant governing
equations, gives a general overview of boundary conditions, and provides a sum-
mary of numerical considerations taken mostly from flow control literature. Then
a brief summary is given for each of the CFD methodologies. Section II.D

This material is declared a work of the U.S. Government and is not subject to copyright protection
in the United States (2009).
*Senior Research Scientist, Computational Aerosciences Branch.

177
178 C. L. RUMSEY AND R. C. SWANSON

describes other CFD models and methods, including blended RANS-LES and
reduced-order models, as well as an in-depth discussion on immersed boundary
methods.
In Sec. III an overview of two specific flow control workshops is provided in
some detail. This description includes additional recent results applied to the same
test cases. Some new flow control strategies—considered from the point of view
of modeling and simulation—are summarized in Sec. IV.

II. Computational Fluid Dynamics Methodologies


In this section, the equations for various approximations of the Navier–Stokes
equations of continuum fluid motion are given. Chemistry and multi-component
gases are not considered, and generally it is assumed that the working fluid is a
perfect gas. There have been a few flow-control studies to date that do not make
these assumptions (see, for example, Damevin and Hoffmann, 2002; Singh and
Roy, 2007)—but perfect gas applications are far more common. We start with the
highest level, DNS, and work toward the lower-level methods which use more
approximations or modeling.
In terms of their representation in Fourier space, the methods described in the
first three sections below—DNS, LES, and RANS—are sketched in Fig. 1. In
these sketches, the energy spectrum is given as a function of wave number k. DNS
resolves all scales of motion; there is no modeling. LES __defines a cutoff wave __
number which is computed directly from the cutoff length D in physical space (D
is typically related to the size of the computational mesh). Wave numbers below
the cutoff are resolved by the simulation, while higher wave numbers (smaller
scales) are modeled. In RANS, turbulence is modeled at all scales.
Throughout this section, the focus is on applications and uses of these method-
ologies for active flow control applications, as found in the literature. However,
the literature presented is only a representative sample: it is by no means an
exhaustive compilation.

A. Direct Numerical Simulation


1. Governing Equations
The Navier–Stokes equations can be written in many different forms. The two
broadest categories are for compressible and incompressible flows. According
to Gad-el-Hak (2000), if radiative heat transfer is neglected, the compressible
equations for a Newtonian isotropic fluid (using Stokes’ hypothesis and neglect-
ing body forces due to external fields such as gravity and electromagnetic
potential) are

∂r ∂ruk
+ =0 (1)
∂t ∂x k

Ê ∂u ∂u ˆ ∂p ∂ È Ê ∂u ∂u 2 ∂u j ˆ ˘
r Á i + uk i ˜ = - + Ím Á i + k - d ki ˙ (2)
Ë ∂t ∂x k ¯ ∂ xi ∂ x k ÎÍ Ë ∂xk ∂xi 3 ∂x j ˜¯ ˙˚
MODELING AND SIMULATION 179

Fig. 1 Simplified representations of three different CFD methodologies in Fourier


space.

2 2
Ê ∂T ∂T ˆ ∂ Ê ∂T ˆ ∂uk 1 Ê ∂ui ∂uk ˆ 2 Ê ∂u j ˆ
rcv Á + uk ˜ = ÁËk ∂x ˜¯ - p ∂x + 2 m ÁË ∂x + ∂x ˜¯ - 3 m Á ∂x ˜ (3)
Ë ∂t ∂x k ¯ ∂x k k k k i Ë j¯

Often, these equations are written in the following strong conservation form:

∂r ∂ruk
+ =0 (4)
∂t ∂x k

∂rui ∂rui uk ∂p ∂ È Ê ∂ui ∂uk 2 ∂u j ˆ ˘ (5)


+ =- + Ím + - d ˙
∂t ∂x k ∂xi ∂xk ÎÍ ÁË ∂xk ∂xi 3 ki ∂x j ˜¯ ˙˚

Ê ˆ
∂r E ∂[( r E + p)uk ] ∂ ÁÁ ∂T È Ê ∂ui ∂uk 2 ∂u j ˆ ˘ ˜˜ (6)
+ = k + Ím + - d ˙u
∂t ∂x k ∂xk ÁÁË ∂xk ÍÎ ÁË ∂xk ∂xi 3 ki ∂x j ˜¯ ˙˚ i ˜˜¯
180 C. L. RUMSEY AND R. C. SWANSON

where the coefficient of thermal conductivity k is often written as k = cpm/Pr,


and the variable E represents the specific total energy E = e + (u2 + v2 + w2)/2.
For a perfect gas, e = cvT, cv = R/(g - 1), g = cp/cv, and the equation of state is
p = rRT, or

[
p = (g - 1) rE - _1_ r(u2 + v2 + w2)
2 ] (7)

In curvilinear coordinates, these equations can be written


∂(G - Gv) _________
∂(F - Fv) _________
∂Q ________
___ ∂(H - Hv)
+ + + =0 (8)
∂t ∂x ∂h ∂z
where Q represents the vector of conserved variables Q = (1/J)[r, ru, rv, rw, rE]T,
J is the Jacobian of the general curvilinear coordinate transformation, J =
∂(x, h, z)/∂(x, y, z), and

È rU ˘
Í ruU + x p ˙
1Í ˙
x

F = Í rvU + x y p ˙ (9)
JÍ ˙
Í r wU + xz p ˙
Í( r E + p)U - x p˙
Î t ˚

È rV ˘
Í ruV + h p ˙
1Í ˙
x

G = Í rvV + hy p ˙ (10)
JÍ ˙
Í rwV + hz p ˙
Í( r E + p)V - h p˙
Î t ˚

È rW ˘
Í ruW + z p ˙
1Í ˙
x

H = Í rvW + z y p ˙ (11)
JÍ ˙
Í rwW + z z p ˙
Í( r E + p)U - z p˙
Î t ˚

È ˘
Í 0 ˙
Í ˙
x t + x yt xy + xzt
Í
Í x xx
˙
xz ˙
Í ˙
1 xt + xt + xt
Í ˙ (12)
Fv = Í x xy y yy z yz ˙
J Í
Í
˙
˙
x t + xyt yz + xzt
Í x xz zz ˙
Í ˙
Í ˙
Í
Î
x f + xy gv + xzh
x v v ˚
˙
MODELING AND SIMULATION 181

È ˘
Í 0 ˙
Í ˙
h t + hyt xy + hzt
Í
Í x xx
˙
xz ˙
Í ˙
1 h t + ht + ht
Í ˙ (13)
Gv = Í x xy
y yy z yz ˙
J Í
Í
˙
˙
h t + hyt yz + hzt
Í x xz zz ˙
Í ˙
Í ˙
Í
Î
h f + hy gv + hzh
x v v ˚
˙

È ˘
Í 0 ˙
Í ˙
z t + z yt xy + z zt
Í
Í x xx
˙
xz ˙
Í ˙
1 zt +zt +zt
Í ˙ (14)
Hv = Í x xy y yy z yz ˙
J Í
Í
˙
˙
z t + z yt yz + z zt
Í x xz zz ˙
Í ˙
Í ˙
Í
Î
z f + z y gv + z zh
x v v ˚
˙

fv = utxx + vtxy + wtxz - kTx (15)

gv = utxy + vtyy + wtyz - kTy (16)

hv = utxz + vtyz + wtzz - kTz (17)


È ˘
Í Ê ∂u ∂u j ˆ 2 ∂uk ˙
t xi x j = ÍÍm Á i + - m d ˙ (18)
Í Ë ∂x j ∂xi ˜¯ 3 ∂xk ij ˙˙
Î ˚

and first-derivative terms with respect to x, y, or z in Eqs. (15–18) are expanded


as fi = xifx + hifh + zifz, where i represents x, y, or z. The contravariant velocities
are given by

U = xt + xxu + xyv + xzw (19)

V = ht + hxu + hyv + hzw (20)

W = zt + zxu + zyv + zzw (21)

For incompressible flows, density is assumed to be constant. As a result, the


energy equation decouples from the continuity and momentum equations.
Furthermore, since in many incompressible applications temperature changes
are insignificant or unimportant, the energy equation is often ignored
completely (Anderson et al., 1984). The incompressible equations for mass and
momentum are

∂uk
=0 (22)
∂x k
182 C. L. RUMSEY AND R. C. SWANSON

∂ui ∂ui uk ∂P ∂ Ê ∂ui ∂uk ˆ


+ =- +n + (23)
∂t ∂x k ∂xi ∂xk ÁË ∂xk ∂xi ˜¯

where the density has been absorbed into the pressure term. For DNS, the incom-
pressible equations, Eqs. (22–23), are often solved instead in vorticity–velocity
formulation (obtained by taking the curl of the momentum equations, which elimi-
nates the pressure terms), where vorticity is defined by w i  -eijk(∂uk/∂xj)

∂w i ∂w i ∂u ∂ Ê ∂w i ˆ
+ uj - wj i = n (24)
∂t ∂x j ∂x j ∂x j ÁË ∂x j ˜¯

The velocity field is obtained from the vorticity via the Poisson equation:

∂ Ê ∂ui ˆ ∂w k
∂x j ÁË ∂x ˜¯ = e ijk ∂x (25)
j j

and the pressure field can be recovered by taking the divergence of Eq. (23) and
simplifying using Eq. (22):

∂2 P ∂ Ê ∂ui uk ˆ
=- (26)
∂xi ∂xi ∂xi ÁË ∂xk ˜¯

There are other ways of writing the Navier–Stokes equations (see, e.g., Zang,
1991; Joslin, 2001), but an exhaustive treatise on this subject is not the intent
here.

2. Boundary Conditions
The boundary conditions at solid walls are typically enforced as no-slip
(u = v = w = 0) with zero normal pressure gradient. For compressible formula-
tions, either adiabatic wall or constant temperature wall conditions are applied.
One difficulty with using the vorticity–velocity formulation is the lack of proper
boundary conditions for the streamwise and spanwise components of vorticity at
the wall. A procedure commonly used is described in Postl and Fasel (2006).
Basically, the numerical method is designed so that these wall values are not
required. The new velocity field is obtained first, then it is used to update vorticity
at the wall, ensuring consistency and zero divergence of the velocity and vorticity
fields. At flow-control boundaries where a transpiration or jet velocity is imposed,
it is common to simply impose a velocity component (e.g., v) or a momentum flux
(e.g., rv), while extrapolating pressure from the interior.
MODELING AND SIMULATION 183

Fig. 2 Sketch of boundary condition locations in a typical flow control simulation


with periodicity imposed at the side planes.

At present, due to computer limitations and the high expense of DNS, all
researchers typically perform computations on a finite span grid with periodic
boundary conditions at the two side planes. The width of the area simulated is
usually limited by the available computational resources. Farfield boundary
conditions vary, although it is common to employ grid stretching, buffer layers, or
special non-reflective boundary conditions to minimize spurious reflections back
into the region of interest. Inflow boundary conditions can play an extremely
important role in simulations: they must generally contain sufficient eddy content
both spatially and temporally to accurately represent the upstream turbulence, or
else a computational inflow length must be provided long enough to allow for
natural development of turbulence prior to the region of interest. A sketch show-
ing the location of typical boundary conditions is given in Fig. 2.

3. Numerical Considerations
By definition, DNS is a computation that resolves all relevant spatial and tem-
poral scales in a flowfield. This means that the grid needs to be fine enough to
resolve features of the order of the Kolmogorov dissipation length scale

h = (n 3/e)1/4 (27)

where n is the kinematic viscosity and e is the dissipation rate. Even with today’s
computers, it is impossible to achieve this resolution at reasonably high Reynolds
184 C. L. RUMSEY AND R. C. SWANSON

numbers. Sandham (2001) cites required near-wall grid spacings for attached
turbulent flow in terms of wall units of the order of Dx+ = 12, Dz+ = 6, and at least
10 points in the wall-normal direction for y+ < 10 to achieve good statistics related
to budgets of Reynolds stresses (these numbers are rules of thumb, and are clearly
scheme dependent). It is estimated that a typical 3-D problem of interest would
require the order of Re(9/4) grid points. See, for example, Joslin (2001), who also
discusses other numerical issues not covered here, such as solving for only
fluctuating components, temporal vs spatial DNS formulations, details concern-
ing disturbance forcing, and coupling with adjoint equations. It is also important
to note that a DNS simulation typically requires long run times. The Kolmogorov
time scale is

t = (n/e)1/2 (28)

Resolving temporal phenomena that occur in this time scale requires time steps
smaller than this scale. The simulation must also go on long enough that the time-
and phase-averaged properties become ergodic (i.e., additional run time and
averaging will not change the result). Guaranteeing this condition can be difficult
to achieve in practice.
It appears to be relatively common to perform “under-resolved” or “coarse-grid”
DNS. In this practice, the full Navier–Stokes equations are solved, but the grid (and
possibly time step) is too coarse to resolve many of the smallest scales of motion.
The argument in favor of this methodology is that the larger resolved scales have
the majority of the influence on most of the relevant aspects of the flowfield. The
downside is two-fold: first, it is difficult to prove the assertion that the smallest
scales are not important for any given problem of interest and, second, calling these
computations DNS can create a false impression for less knowledgeable readers.
Therefore, it is certainly helpful when authors include honest assessments of the
resolution limitations of their simulations. Simulations performed in 2-D, which
are occasionally seen and are also often termed DNS, can sometimes be useful (for
example, to investigate 2-D modes such as nonlinear development of 2-D instabil-
ity waves). However, the loss of all 3-D flow features is often too gross an approxi-
mation for general applications, even when the geometry governing the flowfield is
nominally 2-D. Furthermore, even for statistically 2-D flows, 3-D time-dependent
variation is often required in order to obtain correct 2-D averages.
Recall that the only difference between what is traditionally referred to as a lami-
nar Navier–Stokes simulation and a DNS simulation is resolution. Both methods
solve the same equations, but with laminar Navier–Stokes most small scales remain
unresolved, either because of coarse grid and/or dissipative numerics. Therefore, a
laminar Navier–Stokes computation may be steady (with no time-dependent fea-
tures), but as the grid is refined or as numerics become less dissipative, a time-
dependent computation (with either an inherent or a forced instability mechanism
present in the flowfield) will start to exhibit evidence of time-dependent behavior.
Because this is a continuum process, it is the researcher’s responsibility to demon-
strate the level of resolution achieved. Any simulation exhibiting multiple time-
dependent large- and small-scale features can appear like a DNS simulation, but if
the smallest space and time scales are not resolved, some flow physics are being
omitted and the validity of the simulation remains unclear.
MODELING AND SIMULATION 185

At this point, we will review some of the numerical considerations for DNS
discussed in some flow control papers in the open literature. In Postl and Fasel
(2006), an extra volume forcing term was included in Eq. (24). This term was used
to trip the laminar boundary layer with high-amplitude, time-harmonic 3-D dis-
turbances for selected spanwise Fourier components, to cause the boundary layer
to become turbulent. They also found the spanwise grid extent (in the periodic
direction) to have an influence on their solutions: the larger spanwise extent gave
better results compared with experiment. They used fourth-order compact and
split-compact differences, with a pseudospectral approach with Fourier decompo-
sition in the spanwise direction, and explicit four-stage Runge–Kutta, which is
fourth-order accurate in time. Related work (Laible et al., 2006; Wernz et al.,
2003, 2005) by this group used similar methods, although two of the references
also tried a temporal model.
Lee and Goldstein (2001, 2002) also used the incompressible form with a vor-
ticity–velocity formulation. Their time-stepping methodology is also seen in many
other incompressible papers: it was Adams–Bashforth, with implicit Crank–
Nicolson for the viscous terms (which eliminates the viscous stability constraint).
They employed a “virtual surface” boundary condition, where a localized body
force was used to bring the fluid to a specified velocity. They used the spectral
Chebyshev t method (see Gottlieb and Orszag, 1977) with cosine clustering in the
wall-normal direction.
Rizzetta et al. (1999), Rizzetta and Visbal (2006), and related work by Visbal
and Gordnier (2001) solved the compressible form, Eq. (8), with finite differ-
ences. They also included a vector source term to enforce the geometric conserva-
tion law when performing computations on moving grids. In the earlier work,
their spatial scheme was either second-order central or high-order compact,
whereas in the later work it was a five-point sixth-order compact stencil based on
the pentadiagonal system of Lele (1992), capable of attaining spectral-like resolu-
tion. Also in the later work they employed a tenth-order low-pass Pade-type non-
dispersive filter operator to maintain stability. Temporally they used implicit
approximate factorization (in diagonalized form) with Newton-like subiterations
(typically they only used three subiterations), for second-order accuracy in time.
One important issue they discussed was that of metric evaluations for higher-order
schemes. When using finite differences written in strong conservation form,
metric identities must be satisfied numerically with metric cancellation errors
carefully addressed.
Mittal and co-workers have published several papers in which they directly
solve the incompressible Navier–Stokes equations (Kotapati and Mittal, 2005;
Kotapati et al., 2006, 2007; Mittal et al., 2001; Raju et al., 2007; Ravi et al., 2004;
Utturkar et al., 2003). Of these papers, the first three solved the equations in 3-D;
the others were all 2-D. They employed a second-order accurate central difference
scheme with a two-step fractional-step method (similar to Lee and Goldstein)
for second-order time advancement. The method was cell-centered, collocated
(non-staggered), where they computed both the cell-center velocities and the face-
center normal components. The 3-D simulations were for jet flowfields. They
found it helpful to introduce three-dimensionality into their solutions via an initial
sinusoidal spatial perturbation on the flowfield. Jet boundary conditions were
applied at the bottom of a jet plenum included in the computations. In some of the
186 C. L. RUMSEY AND R. C. SWANSON

2-D work, they modeled the diaphragm within the plenum as a moving body,
rather than with an imposed velocity boundary condition.
Deng et al. (2007) used the compressible form of Eq. (8) to solve separated flow
over an airfoil with steady and pulsed jet control. Similar to Rizzetta, they used
sixth-order compact differencing based on the method of Lele, as well as a high-
order compact filter. They used LU-SGS with three sweeps through the mesh,
along i + j + k = constant planes. The left-hand side used first-order upwinding.
They did not find it necessary to seed disturbances into the flowfield, because for
their problem the instability wave in the separated shear layer was unstable enough
to trigger turbulence.
Barwolff et al. (1996), Wengle et al. (2001), and Neumann and Wengle (2003)
performed thorough DNS simulations over a controlled backstep flow, in which
they attempted to resolve all the relevant scales in the turbulence field. The Reynolds
number based on step height was only 3000, and the oncoming boundary layer had
Req = 285. In their latter paper the DNS grid had over 47 million cells. They solved
the incompressible equations on a staggered, non-uniform, Cartesian grid. They
employed both a fourth-order compact scheme and a second-order central scheme
for some runs. The temporal advancement used an explicit second-order leapfrog
scheme with time-lagged diffusion. In the earlier papers, the boundary layer was
laminar and the flow transitioned on its own to turbulent in the separated shear
layer. In the latter paper, to induce fully-developed turbulence in the boundary
layer, they used a series of vorticity generators (modeled as blocked-out surface-
mounted thin vertical fins) near the inlet. They found it extremely important in
general to generate the same inflow conditions in the simulation as in the experi-
ment. For the case in which harmonic blowing/suction was applied, they employed
time-dependent boundary conditions directly on the wall: the u and w components
of velocity were specified on two neighboring crosswind rows of gridpoints.
Sumitani and Kasagi (1995) performed DNS for channel flow with uniform
wall injection and suction. The flow was at low Reynolds number, Ret = 150. They
solved a fourth-order PDE for velocity, a second-order PDE for the wall-normal
component of vorticity, and the continuity equation. A spectral method (with up to
128 by 128 Fourier modes in wave number space) was used in x and z, and a
Chebychev polynomial expansion, up to 96th order, was used in y, the wall-normal
direction. They employed periodic boundary conditions in both x and z. The col-
location grid used to compute the nonlinear terms in physical space had a factor
of 1.5 times finer resolution in each direction in order to remove aliasing errors.
Similar to other work discussed above, the time integration was second-order
Adams–Bashforth for the nonlinear terms and Crank–Nicolson for the viscous
terms. For injection and suction, the wall normal velocity component was imposed
at the walls. One of their checks was that no energy accumulation occurred at high
wave numbers in the energy spectra; such accumulation would be evidence of
insufficient numerical resolution.

4. Summary, Issues, and Limitations


To summarize, DNS is a very powerful tool and can be used successfully for
flow control simulations. Currently, use of true DNS is limited to very low
MODELING AND SIMULATION 187

Reynolds numbers, but some researchers claim that “under-resolved” DNS at


higher Reynolds numbers can also be useful for predicting certain quantities that
do not appear to depend on resolving the smaller scales in the flow. This type of
validation for active flow control problems remains on a case-by-case basis, as it
relies primarily on comparison with experiment. Arguably, “under-resolved” DNS
may be characterized as an implicit LES technique, to be discussed in the next
section. DNS often requires the introduction of instabilities into the simulation to
ensure that turbulence develops where desired; there appear to be many accept-
able methods for doing this.
Clearly, low-dissipation schemes are required for DNS. Many methods seem to
be successfully employed, including second-order central differencing, higher-
order compact differencing, and pseudo-spectral/spectral methods. The spatial
order of accuracy of the scheme influences grid requirements, but we do not dis-
cuss this aspect here. It is also important to recognize that additional dissipation is
required to stabilize some numerical schemes. For example, compact schemes are
often stabilized through the use of additional filtering. As discussed by Gaitonde
et al. (1999), the order of the filter can have a significant impact: lowering the
order provides better filtering at the spurious frequencies, but at the cost of reduc-
ing the range of resolvable frequencies.

B. Large Eddy Simulation


1. Governing Equations and Turbulence Modeling
Sagaut (2006) stated: “In practice, the Large-Eddy Simulation technique
consists of solving the set of ad hoc governing equations on a computational
grid which is too coarse to represent the smallest physical scales.” In other
words, in LES the dynamics of the large-scale structures in the flowfield are
computed, whereas the effect of small-scale turbulence is modeled or neglected.
Formally, the governing equations are derived by applying a low-pass filter with
non-uniform filter width to the Navier–Stokes equations (Vasilyev, 2001). As
discussed in Sagaut, the LES low-pass filter can be defined as a convolution
product:

+• +•
3
f ( x, t ) = Ú Ú f (x , t ¢)G( x - x , t - t ¢) dt ¢ d x
-• -•
(29)

__ __
where f (x , t) is the resolved part of a space–time variable f (x , t). By assuming the
filter is commutable, the filtered governing equations can be obtained [discussion
on the errors associated with this assumption can be found in Geurts (1999)].
For the compressible Navier–Stokes equations, the equations contain both
ordinary and Favre-filtered variables (see, e.g., Knight et al., 2001). The Favre
filter is defined by

rf
f = (30)
r
188 C. L. RUMSEY AND R. C. SWANSON

where the overbar denotes the ordinary spatially filtered variable, and the tilde
denotes the Favre-filtered variable. After filtering, the compressible governing
equations [Eqs. (4–6)] become

∂r ∂r u k
+ =0 (31)
∂t ∂x k

∂r u i ∂r u iu k ∂p ∂T ik
+ =- - (32)
∂t ∂x k ∂xi ∂x k

∂r E ∂[( r E + p )u k ] ∂Hk


+ = (33)
∂t ∂x k ∂x k
__ __
where the total stress tensor is Tik  rtik - sik, and the sum of the heat flux plus
work Hk  Qk + k ∂T̃/∂xk + Tikui with k often expressed as cpm/Pr, and

Ê ∂u ∂u 2 ∂u j ˆ


s ik = m Á i + k - d ik ˜ (34)
Ë ∂xk ∂xi 3 ∂x j ¯

As a subtle point, it should be noted that as a result of the filtering, the effect of
the filtered stress kinetic energy (k  tii/2) should
__ be __ included
__ in the definition__
of the total energy (Knight et al., 2001), i.e., rẼ = rẽ + r(ũ 2 + ṽ2 + w̃2)/2 + rk.
Then, the equation of state becomes

È 1 ˘
p = (g - 1) Ír E - r (u 2 + v 2 + w 2) - r k ˙ (35)
Î 2 ˚

However, for lower speed flows, the effect of k is usually negligible, and many
researchers ignore it in the definition of energy and equation of state.
As a result of the filtering, two unknown quantities emerge, the subgrid scale
stress tik and the heat flux Qk:

t ik = u  iu k
i uk - u (36)

Qk = - c p r ( u  kT )
kT - u
(37)

Note that some researchers define tik with the opposite sign, and also that tik some-
times includes the density term in its definition. These differences do not matter
as long as everything is carried through consistently. The quantities in Eqs. (36)
and (37) must be modeled. A discussion on the many models proposed and used
is beyond the scope of this chapter; only a few are briefly outlined at the end of
this section. We recommend Sagaut (2006) for details and many other references.
In general, as discussed in Sagaut, there are two basic methods: explicit and
implicit modeling. In explicit modeling, a subgrid model is explicitly introduced.
MODELING AND SIMULATION 189

This model may be based on functional modeling, which attempts to represent the
nature of the interscale interactions and energy transfers, or on structural model-
ing, based on mathematical expansions. In implicit modeling (also known as
ILES), no extra terms are introduced into the governing equations, but the numeri-
cal method is selected such that the numerical error fulfils desired properties and
effectively acts like a subgrid model. Implicit modeling effectively assumes that
the action of subgrid scales on the resolved scales is strictly dissipative. With no
explicit model present, ILES appears to be functionally similar (from an imple-
mentation point of view) to “under-resolved” DNS; neither method resolves the
finest scales, and both require enough inherent dissipation in the numerical scheme
to prevent the non-physical build-up of energy at the smallest resolved scales.
For incompressible flow, the form of the filtered equations is simplified.
Equations (22) and (23) become

∂uk (38)
=0
∂x k

∂ui ∂ui uk ∂P ∂ Ê ∂ui ∂uk ˆ ∂t ik (39)


+ =- +n + -
∂t ∂x k ∂xi ∂xk ÁË ∂xk ∂xi ˜¯ ∂xk

where

t ik = ui uk - ui uk (40)

Note again that some researchers define tik with the opposite sign. Also note that
in Eq. (39) the density has been absorbed into the pressure term. Here in the
incompressible equations, the subgrid scale stress tik is unknown and must be
modeled.
The most commonly used subgrid model is arguably the eddy viscosity type,
which can be written as

Ê 1 ˆ 2
t ik = - 2n t Á S ik - d ik S jj˜ + d ik k (41)
Ë 3 ¯ 3

where the second term in the parentheses is zero for incompressible flows, and
the effects of the filtered stress kinetic energy (k  tii/2)_ in the last term are
often ignored, particularly for lower-speed flows. The term Sik is the local resolved
rate of strain:

1 Ê ∂u i ∂u k ˆ
S ik = + compressible (42)
2 ÁË ∂xk ∂xi ˜¯

1 Ê ∂ui ∂uk ˆ
= + incompressible (43)
2 ÁË ∂xk ∂xi ˜¯
190 C. L. RUMSEY AND R. C. SWANSON

The commonly used Smagorinsky model for the eddy viscosity term is

2
n t = ( cs D ) | S | (44)
_ _____
__ __
with | S | = ÷2__Sij Sij , and D typically defined by some measure of the local grid spac-
ing, such as D = (D x Dy Dz)(1/3). The cs term is a constant for the original Smagorinsky
model, commonly set to around 0.1.
Note that in most subgrid models, neither the formal filter function nor the filter
width are explicitly defined. Instead, the models have__a built-in filter, related to the
local grid spacing: for the Smagorinsky model it is cs D. This built-in filter controls
the size of the smallest locally resolved flow structures. If the filter width is
increased (for example by increasing cs), the solution becomes smoother because
of increasing diffusion, but the modeling error increases (Brandt, 2007). It can be
difficult to choose the optimum filter width, which has been noted to be flow
dependent. This difficulty has been one reason for the development and subse-
quent success of the widely used dynamic Smagorinsky model of Germano et al.
(1991). This model dynamically computes a variable cs term, rather than setting it
to a constant value.
One-equation transport models are also sometimes solved for k, in which case
the eddy viscosity term is commonly approximated by

n t = cm k (1/2) D (45)

Models such as the ones described above generally work well when the propor-
tion of shear stress carried in the subgrid model is very small compared to that
carried in the resolved large eddies (Sandham, 2001).
For compressible flows, an eddy viscosity model for the subgrid-scale heat flux
is (see Urbin and Knight, 2001):

rc pn t ∂T
Qk = (46)
Pr t ∂xk

2. Boundary Conditions
The boundary conditions for LES are generally the same as for DNS. In particu-
lar, inflow boundary conditions can still play a crucial role in determining the
success of a simulation. Additionally, if any subgrid scale model employs
transport equations, then boundary conditions must be defined for the relevant
variables. Also, although LES computations have become more affordable over
recent years due to increased computational capabilities, most researchers still
perform simulations on a finite span grid with periodic boundary conditions at the
two side planes, with the width of the area typically limited by the available
computational resources.
Approximate boundary conditions are sometimes applied for LES at solid
walls, to mitigate the need for fine grid resolution near the wall. These wall models
are designed to approximate the effects of the important wall layer dynamics, and
MODELING AND SIMULATION 191

to capture the log-law behavior of turbulent boundary layers. They have been
applied to flows with wall injection and suction by Piomelli et al. (1989).

3. Numerical Considerations
LES is in some ways more difficult than DNS, in the sense that it is hard to
know and/or prove that a simulation is “good enough.” Not only is the solution
influenced by the choice of subgrid model (or lack thereof in the case of ILES),
but it is also strongly dependent on choice of numerical algorithm and grid.
Furthermore, as discussed in Sagaut (2006), there are three categories of scales—
subgrid scales, subfilter scales, and physically resolved scales—and the relative
relationships between these three play an important role. Unlike the RANS meth-
ods to be discussed later, it is very difficult and expensive to determine the numeri-
cal accuracy of an LES simulation based on grid density influence. Most
researchers perform only one simulation for a given case, typically at or near the
maximum grid resolution that they can afford at the time. Also, LES suffers from
the same difficulty as DNS in that it is often very costly to run a simulation for a
long enough time for statistical quantities to be guaranteed temporally converged
in their temporal or phase-averaged mean.
There is some guidance available to estimate the number of grid points needed
for LES at a given Reynolds number. When the filter cutoff is in the inertial range
of the energy spectrum, the resolution required by LES is weakly dependent on
the Reynolds number. Based upon the analysis of Chapman (1979), discussed by
Piomelli and Balaras (2002), the number of points required in the outer part of a
developing boundary layer (i.e., approximately 90% of the boundary layer) is
proportional to Re0.4. Chapman also points out that if the inner part of the bound-
ary layer (i.e., viscous sublayer) is resolved, which is sometimes called highly
resolved LES, the number of points required for this region is in the order of Re1.8.
By resolving the viscous sublayer, the resolution requirement approaches that of
DNS, making LES impractical for high Re flows. Only by modeling the subgrid
scales of the turbulent motion do we retain the weak mesh dependence on Reynolds
number. The modeling bears the burden of reproducing the energy transfer of the
small scales and, thus, determining the ultimate success of the simulation.
Verification of flow control results from LES is an especially important require-
ment if the objective is to use the results from LES as data to enable improved
turbulence modeling. As pointed out by Mason (1994), a key test for verification
is to demonstrate convergence of the results as the numerical resolution is increased
and the filter scale is reduced. Mason suggests that a credible test of convergence
ought to involve simulations spanning at least a factor of four in resolution.
Certainly, the highest resolution case depends upon available resources, and the
lowest resolution case must be a realistic simulation. Mason also points out that
this convergence test may not be sufficient due to boundary regions (e.g., solid
boundaries) where at high Reynolds numbers the subfilter model may dominate.
In general, supplemental tests in which the LES results are compared with detailed
experimental data are also necessary to confirm the behavior of the solution.
Before discussing flow-control related applications of LES, it is worth mention-
ing some of the lessons learned from Sagaut (2006), since many of these lessons
apply to flow control problems as well: a) for simple (ideal) cases, many LES
192 C. L. RUMSEY AND R. C. SWANSON

models and methods work well, but extending to inhomogeneous cases can be
difficult for subgrid models and numerical methods; b) shear flows demonstrate
strong sensitivity to inflow boundary conditions (see also Sagaut and Le, 1997),
and it is not known how best to generate these conditions; c) low numerical error
and consistent modeling are important in regions near the wall, transition areas, or
other areas where “flow driving” mechanisms are present, but are not as important
in regions where the energy cascade is the dominant mechanism; d) explicitly
applied subgrid models can provide dissipation and hence can stabilize complex
simulations that have a tendency to go unstable; e) there is a consensus today that
the numerical accuracy must be at least second order in space and time, although
many researchers use higher order than second in space; f) LES is most useful for
massively separated industrial flows, where the large scales are not driven by
dynamical details of the boundary layers, but for fully attached flows LES can be
problematic and is still too expensive for high Reynolds number simulation on a
routine basis. This latter point is one of the reasons for the development and use
of approximate boundary conditions for LES (see, e.g., Piomelli et al., 1989).
Dandois et al. (2006a, b) performed compressible LES computations for syn-
thetic jet flows. They used a selective mixed-scale model (a Boussinesq-like
approximation to the subgrid scale stress tensor). The spatial scheme was based
on the AUSM + (P) scheme, along with second-order central differences for the
viscous fluxes. Implicit second-order accurate Gear’s backward time scheme was
employed, with the inversion of the linear system at each iteration done by the LU
symmetric Gauss–Seidel implicit method. Enough subiterations were performed
to achieve one order reduction in the residuals: typically the number needed was
8. In order to achieve a turbulent boundary layer, they based their inflow boundary
conditions on a steady RANS profile, with turbulent fluctuations superimposed.
The grid used had cell spacings in the refined zone near the jet of approximately
Dx+ = 50, Dy+ = 20, and Dz+ = 1, where z was the direction normal to the wall.
They included the plenum flow in their simulations, and imposed a sinusoidal
velocity boundary condition at the bottom wall of the plenum.
Slomski et al. (2006) used a compressible solver with a subgrid scale eddy
viscosity model formed from a transport equation for the subgrid scale turbulent
kinetic energy. They used a fifth-order spatially accurate upwind-biased scheme
for the nonlinear convection terms that allowed the reduction of dissipation inher-
ent in the upwind formulation. Their flowfield, a jet around a curved airfoil surface,
included the jet plenum, and a mass flow was specified as the boundary condition
for the jet. A 2-D RANS solution was used as an initial condition for the simula-
tion, and the solution was allowed to transition to turbulence on its own. Their grid
had cell spacings of approximately Dx + = 20 - 320, Dz+ = 30, and Dh+ = 1, where
h was the direction normal to the wall and z was spanwise.
Yuan et al. (1999) solved the incompressible form of the equations for a round
jet in crossflow. They used a dynamic subgrid model, and a fractional step, non-
staggered solution technique. The semi-implicit time advancement scheme was
second order. They defined the grid size volumes next to the jet exit to be small
enough to ensure that virtually all scales of motion were resolved near the solid
surface (Dx+ = 1.4, Dz+ = 2.2). Therefore, in this region, the subgrid scale contri-
butions to the turbulent stresses were negligible. They found particular sensitivity
of the flowfield to characteristics of the inflow jet: they tried several methods and
MODELING AND SIMULATION 193

found plug flow (not including any of the pipe in the simulation, but specifying
velocity profiles directly on the plate surface) to be the worst compared to
experiment. For turbulent pipe inflow, profiles were obtained from auxiliary
simulations.
Dejoan and Leschziner (2004, 2005) solved the incompressible form of the
equations for a perturbed back step and a plane wall jet. The method was second
order in space, with central differencing for both advection and diffusion. A
second-order fractional-step time marching method with a backward-biased
approximation was employed. Flux terms were advanced explicitly with Adams–
Bashforth. The pressure was computed from the pressure-Poisson problem by
partial diagonalization and multigrid in conjunction with successive line over-
relaxation. Several different subgrid scale models were tried, including
Smagorinsky and dynamic Smagorinsky. For the back step, the inlet conditions
were created from a precursor simulation for fully developed channel flow, and
the resulting time-dependent realizations were fed into the back step at the inflow
boundary. For the jet flow, the profile was generated from experiment, and
included random isotropic fluctuations with variance consistent with the experi-
mental turbulence level. They did not include any plenum in their computations:
surface boundary conditions were applied at the wall. For grid size, the spacings
for the back step had D x + = 28, Dy+ = 1.5 (wall) 4.5 (shear layer), and Dz+ = 20,
__ D x = 24, __Dy < 1, and Dz = 23.
where z was the spanwise direction; for the jet + + +

An interesting statistic that they plotted was D /h, where D = (D x Dy Dz)(1/3) rep-
resented the grid size, and h was the Kolmogorov length scale, Eq. (27). (To
determine h they had to compute the balance of the turbulence energy budget;
the e term could not be obtained explicitly because__a fraction of it was contained
in the subgrid scales.) They showed that the ratio D /h was everywhere less than
about 10, indicating that the cut-off for their simulations was close to the dissipa-
tive part of the wave-number range.
Chang et al. (2002) solved incompressible flow in a channel to explore the
concept of opposition control, using a dynamic Smagorinsky model. Opposition
control uses distributed suction and blowing to oppose the motion of near-wall
turbulent structures. They employed a hybrid Fourier-spectral and second-order
finite difference method: the Fourier-spectral method was used to compute the
spatial derivatives in homogeneous directions, and the finite difference method
was used in the wall-normal direction on a staggered grid. Crank–Nicolson was
used for wall-normal derivatives, and an explicit third-order Runge–Kutta scheme
was used for terms involving derivatives in the homogeneous directions. Flow-
control boundary conditions were applied at the wall, and wall spacings were
Dx+ = 40 - 70, Dy+ < 1, and Dz+ = 11 - 24.
Rizzetta and Visbal (2003b) solved cavity flow with the compressible LES
formulation. Simulations were done at a somewhat lower Reynolds number than
experiment, and a dynamic Smagorinsky model was used. Their numerical algo-
rithm was similar to that described earlier for their DNS applications (Rizzetta
and Visbal, 2006; Rizzetta et al., 1999; Visbal and Gordnier, 2001). For their inflow
boundary condition, they ran an auxiliary flat plate simulation with perturbed
variables to generate turbulence, using no subgrid model. After the auxiliary simu-
lation was recorded and forced to be periodic in time, it was used as a database of
5000 profiles for time-dependent inflow boundary conditions on the cavity flow
194 C. L. RUMSEY AND R. C. SWANSON

problem. They used an initial condition from a 2-D RANS solution, and they did
not include any flow-control plenum in their computations (the jet velocity was
specified directly on the wall). For the solid wall boundary condition, in addition
to no-slip they prescribed a fourth-order accurate representation of the zero nor-
mal pressure gradient. Wall spacings were: Dx+ = 6 (near lip) 33 (near center of
cavity), Dy+ = 1.6, and Dz+ = 7.5.
You et al. (2006b) and You and Moin (2006) solved the incompressible equa-
tions, and employed a dynamic Smagorinsky-type eddy viscosity model. They
used a nondissipative second-order central-difference spatial algorithm and a fully
implicit fractional step method that avoids severe time-step restrictions, based on
numerical stability constraints. All terms including diffusion terms used Crank–
Nicolson, with a Newton iterative method for solving the discretized nonlinear
equations. The Poisson equation was solved by a hybrid procedure, combining
multigrid for the curvilinear planes and a Fourier spectral method for the remain-
ing Cartesian direction. In a hump case they did not model the flow-control ple-
num, and in an airfoil case they did. In the hump case, turbulent inflow profiles
were provided from a separate simulation of a flat plate boundary layer. Wall
spacings were: Dx+ = 50, Dy+ < 1, and Dz+ = 25 for the hump and D x+ = 60,
Dy+ = 1.2, and Dz+ = 16 for the airfoil.
Other flow-control LES papers (Jones and Wille, 1996; Kjellgren et al., 2000;
Lesbros et al., 2006; Neumann and Wengle, 2001; Suponitsky et al., 2005) are not
discussed here. All used the incompressible form of the equations, and all included
either a Smagorinsky or dynamic Smagorinsky model.

4. Summary, Issues, and Limitations


To summarize, LES is now being used by many researchers for flow control
type problems, although not on a routine basis for industrial applications. Most
have been employing explicit subgrid models: usually of the dynamic Smagorinsky
type. Although high-order algorithms are being used by many researchers, non-
dissipative second-order spatial algorithms (central differencing) are certainly
still very common, and appear to be acceptable for these types of problems. Issues
related to order of accuracy, dissipation, and filtering—as discussed earlier in the
DNS section—obviously apply to LES as well. Also, like DNS, there appear to be
many acceptable methods in LES for introducing instabilities (when needed) into
the simulation to ensure that turbulence develops.
As mentioned in the literature review, some researchers model a flow control
plenum or jet pipe and some apply blowing/suction boundary conditions directly
on the wall surface. The plenum is usually modeled when there is concern about
capturing details of the flow interactions near the plenum exit. Regarding grid
spacing in terms of wall units, most researchers tend to employ spacings of the
order of Dz+ = 20 in the spanwise (periodic) direction, and Dy+ = 1 at the wall in
the normal direction, with grid stretching away from the wall. In the streamwise
direction the variation among researchers is larger, but numbers of the order of
Dx+ = 50 are typical. The requirements may vary depending on the application.
For example, Yuan et al. (1999) used DNS-type wall spacing in what they believed
to be a particularly sensitive region of the flowfield, to avoid subgrid scale model-
ing in that area.
MODELING AND SIMULATION 195

C. Reynolds-Averaged Navier–Stokes
1. Governing Equations and Turbulence Modeling
In the well known Reynolds decomposition, the flow variables in the Navier–
Stokes equations are decomposed into mean and fluctuating components:

f = f + f¢ (47)
__
The average of a fluctuating quantity is zero, f ¢ = 0. For compressible flows, Favre
averaging is used, as defined in Eq. (30), with the exception that the overbar now
denotes a mean variable rather than a spatially filtered variable. The RANS equa-
tions turn out to be identical to the spatially filtered equations used for LES. That is

∂r ∂r u k
+ =0 (48)
∂t ∂x k

∂r u i ∂r u iu k ∂p ∂T ik
+ =- - (49)
∂t ∂x k ∂xi ∂x k

∂r E ∂[( r E + p )u k ] ∂Hk


+ = (50)
∂t ∂x k ∂x k
__ __
where the total stress tensor is Tik  rtik - ~sik, and the sum of the heat flux plus
work done by__ the stresses is Hk  Qk + k ∂T/∂xk + Tikũi with k often expressed as
cpm/Pr, and sik is given by Eq. (34).
For incompressible flow, the RANS equations are also identical to the spatially
filtered LES equations:

∂uk
=0 (51)
∂x k

∂ui ∂ui uk ∂P ∂ Ê ∂ui ∂uk ˆ ∂t ik


+ =- +n + - (52)
∂t ∂x k ∂xi ∂xk ÁË ∂xk ∂xi ˜¯ ∂xk

Just as for LES, the unknowns tik (and Qk for compressible flow) must be mod-
eled. For RANS, Qk is probably most often modeled as

rc pn t ∂T
Qk = (53)
Pr t ∂xk

The turbulent stress tik can be modeled in many ways. These methods include the
high-level second-moment closure modeling (full Reynolds stress modeling),
where a transport equation is solved for the turbulent dissipation rate e (or equi-
valent quantity) as well as for each stress component (Hanjalic and Jakirlic,
2002; Pope, 2000; Wilcox, 2006). They also include the more commonly used
196 C. L. RUMSEY AND R. C. SWANSON

linear and nonlinear eddy viscosity models, for which the turbulent stress can be
written as
Ê 1 ˆ 2
t ik = - 2n t Á S ik - d ik S jj˜ + d ik k + Fik (54)
Ë 3 ¯ 3

For linear models, Fik = 0 and Eq. (54) is identical to Eq. (41). For nonlinear and
explicit algebraic stress models, Fik is a function of various tensor bases, depend-
ing on the model (see Gatski and Rumsey, 2002, for more details).
Although the RANS equations are identical in form to the filtered LES equa-
tions up to the point where tik and Qk appear as extra terms in the momentum and
energy equations, the models used are very different. For example, in the RANS
equations the models for determining the eddy viscosity term nt do not depend
__ on
grid parameters. In LES, because of the filtering operation, a term like D, related
to local grid cell size, is involved. Most widely used RANS turbulence models
typically involve the solution of one or more additional transport equations, which
are active everywhere and have an influence across the entire energy spectrum.
For example, the Spalart–Allmaras model (Spalart and Allmaras, 1994) solves
one equation for a variable directly related to nt. The k-w model of Wilcox (2006)
and the k-w SST model of Menter (1994) solve two equations for the turbulent
kinetic energy k and the dissipation per unit turbulent kinetic energy w, and these
determine eddy viscosity via
k
nt = (55)

where w is either w or another function based on a stress limiter to improve per-
formance for separated flows. The k-e family of models solves two equations for
the turbulent kinetic energy k and the turbulent dissipation rate e, and these deter-
mine eddy viscosity via
k2 (56)
n t = cm
e

2. Boundary Conditions
The boundary conditions for RANS are generally the same as for DNS and LES.
However, the solution of transport equations for turbulence quantities involves
additional boundary conditions on the turbulence terms. At solid walls, k = 0, but
the boundary conditions on w and e, for example, are not as straightforward. It is
known how these variables behave as they approach the wall (Wilcox, 2006), but
there are many opinions about the best way to impose their boundary conditions at
the wall. A method commonly used for w is described in Menter (1994):
6n
w w = 10 (57)
b1d 2

where b1 is one of the constants in the k-w model, and d is the distance to the next
point away from the wall.
MODELING AND SIMULATION 197

In order to accurately compute turbulent flows near solid walls, the wall-normal
grid spacing needs to be fine enough that the first grid point lies near or within a
distance d+ = 1 for most turbulence models. However, similarly to LES, approxi-
mate boundary conditions are sometimes applied at walls when using RANS.
These so-called wall functions relax the restriction on required wall spacing, but
are strictly valid only for attached boundary layers.
At flow-control boundaries where a transpiration or jet velocity is imposed, it is
unclear how best to handle the turbulence terms when the flow is entering the
domain. If the inflow is supposed to be turbulent, then some imposed boundary
conditions representing turbulence should be employed. This requirement is one
of the many reasons why researchers sometimes choose to model a plenum region
inside a transpiration surface. Then, turbulence can develop naturally prior to
emerging from the plenum, and (hopefully) give desired turbulence levels. An
example is shown from a computation of flow emanating from a blowing slot on
the upper surface of an airfoil in Fig. 3. Here, the computed turbulence levels
(eddy viscosity) are shown as gray-scale contours behind the mean flow velocity
vectors. The turbulence has been allowed to develop on its own through the ple-
num leading to the slot exit. Because of the shape of the plenum (concave curvature
on upper wall, convex curvature on lower wall), the turbulence is stronger near the
upper wall of the slot than near the lower wall.

Fig. 3 Flow near the upper-surface blowing slot of an airfoil, showing normalized nt
contours along with mean flow velocity vectors (the lower channel represents the slot,
with the plenum to the left).
198 C. L. RUMSEY AND R. C. SWANSON

3. Numerical Considerations
Once the determination has been made that the approximations inherent in the
RANS equations are well founded for flow control applications, the main numeri-
cal/modeling considerations are the numerical accuracy (i.e., adequacy of grid
and time step), and choice of turbulence model. The former is usually determined
through a sequence of grid and time step studies; in fact, most refereed journals
require this sort of numerical analysis for computed submissions. The latter is
often a matter of validation against experiment. This validation can be difficult,
since it is possible that certain turbulence models will work well for some situa-
tions or configurations and poorly for others. Also, it is hard to guarantee that all
of the boundary conditions match experiment, as will be discussed below.
One advantage of RANS over DNS and LES is that 2-D computations are fully
justified options. This is because in RANS the large-scale 3-D structures are not
resolved, but rather their effects are modeled in an average sense. Thus, RANS
computations can be a very inexpensive way to explore the flow physics of many
nominally 2-D applications. However, it can be very difficult to conduct an experi-
ment that is sufficiently 2-D, particularly when the flowfield is separated or inher-
ently unsteady (common for flow-control applications).
There have been many papers for RANS applications to flow control problems.
In the interest of space, here we only mention some of the ones we found in the
refereed literature, in order to discuss some of the numerical considerations
brought up by the authors. Some additional papers are mentioned in Sec. III on
CFD validation within the context of specific flow control validation cases. The
interested reader can also find many additional examples, particularly among
AIAA conference papers and journals.
Three works from the early 1990s used the compressible RANS equations to
compute 3-D jets in crossflow (Claus and Vanka, 1992; Demuren, 1993; Kim and
Benson, 1992). They used various turbulence models, including k-e, multiple-
time-scale, and full Reynolds stress. At the time, a 256 × 96 × 96 grid (2.4 million
points) was considered extremely fine, but even then it was recognized that this
level was not free from noticeable discretization errors. One author questioned the
use of RANS for flows with inherent large-scale structures. Another found it
necessary to include a part of the jet-emitting pipe in the computation, rather than
applying a jet profile directly on the wall, while the third author specified the jet
boundary conditions on the wall by using experimental data. These conditions
were difficult to specify because of the crossflow interaction, and it was noted that
they “are rarely measured in sufficient detail” in experiments (Demuren, 1993).
Another work from the 1990s was Wu et al. (1998), who computed 2-D compres-
sible RANS for post-stall airfoil control. They used second-order central differenc-
ing with fourth-difference artificial dissipation, and a second-order implicit scheme
in time. Blowing and suction were applied on the surface of the airfoil. Their grids
were surprisingly coarse: the finest was only 181 × 261. A simple algebraic turbu-
lence model was used, despite the fact that it was known to overestimate the lift for
separated flow. However, other more advanced turbulence models they attempted
to use yielded puzzling or inconsistent behavior for their problem.
Ekaterinaris (2004) used 2-D incompressible RANS with a pseudo-
compressibility correction for active flow control on an airfoil. He used third-order
MODELING AND SIMULATION 199

upwind-biased convective fluxes along with second-order central differencing for


the viscous fluxes, and noted that use of higher order is not guaranteed to be better
for stretched meshes unless all other quantities, including metrics, are also evalu-
ated at higher order. The Spalart–Allmaras model was used, and flow control was
applied directly at the jet-exit surface (no plenum). From previous investigations,
he noted that coupling the turbulence model with the mean flow equations—as
opposed to solving it uncoupled—made little difference. He also noted that fine
grid resolution was needed near the jet port.
Gross and Fasel (2006a) solved both 2-D and 3-D compressible RANS over a
blown cylinder, using fifth-order upwind WENO along with fourth-order for the
viscous terms. They used several different turbulence models, which were solved
spatially second order. Flow was prescribed at the nozzle exit (no plenum). They
found differences with experiment for all models run in 2-D. But, in addition to
the models themselves, there were other potential sources for error noted: a)
strong dependence on the nozzle boundary condition, which was not documented
in the experiment; b) the flow was transitional, a condition that the turbulence
models were not designed to handle; c) in reality there were 3-D structures present
that the turbulent mean flow could support and amplify, as demonstrated in 3-D
calculations.
Rehman and Kontis (2006) used incompressible 2-D RANS in a segregated
commercial flow solver for an airfoil with flow control. Six different turbulence
models were attempted. It was not entirely clear from their paper whether they
applied the flow control boundary condition on the airfoil surface or inside a
plenum. They concluded that the k-w SST model of Menter worked best for their
cases. Gustafsson and Johansson (2003) also used the same flow solver, but for a
3-D application of jet in crossflow. They included the plenum in the computation,
and concluded that a k-e model was poor, the SST model was reasonably good,
and a full Reynolds stress model was best compared with experiment.
Guo et al. (2003) used incompressible 2-D RANS to explore synthetic jet vec-
toring. Upwind fifth-order spatial differencing for convective terms was combined
with second-order central differencing for viscous terms, and the method of
pseudocompressibility was used with subiterations to advance in time. The
Spalart–Allmaras model was employed, and a plenum was included in the com-
putations. At the bottom of the plenum, time-dependent velocity boundary condi-
tions were applied. The time-harmonic normal velocity perturbation was taken
into account at the synthetic jet boundary, yielding a modified boundary condition
for the pressure.

4. Summary, Issues, and Limitations


It is difficult to characterize the use of RANS for flow control applications,
because there has been such a wide range of applications to date. Certainly, it
appears that turbulence modeling comes under question quite often for its appli-
cability for these flows. One concern is that RANS may not be strictly valid for
unsteady flows. But when run in unsteady mode (often referred to as URANS),
one can argue that if the time scale of any gross unsteady motion is much greater
than the physical time step employed, which in turn is much greater than the time
scales associated with the turbulence, then the use of a RANS turbulence model is
200 C. L. RUMSEY AND R. C. SWANSON

justified. Nonetheless, the fact that turbulence models have been designed and
calibrated based on steady flows leaves room for uncertainty.
Overall, RANS applications to flow control problems have been mixed. There
have been many successes, but also many areas where performance has been fair
or poor. In large part, success or failure may depend entirely on the particular
problem and quantity of interest; for example, in certain cases RANS may miss
absolute lift or drag levels, but capture the increments due to changes in flow
control parameters.
Although turbulence modeling is certainly a possible cause for poor RANS
results, it may not always be the primary reason. It is usually difficult to simulate
precisely the same conditions as experiment. In particular, for flow control prob-
lems the boundary conditions at flow control interfaces are often not defined in
sufficient detail to make the assignment of CFD boundary conditions entirely
clear. Furthermore, conducting experiments in unsteady flow control can be much
more difficult than in steady problems, and measurement uncertainties can be
larger (Rumsey et al., 2006).
Numerically, many different schemes have been used for solving the RANS
equations. Issues discussed earlier in the DNS and LES sections related to order
of accuracy, dissipation, and filtering apply here as well. It is safe to say that
although higher-order methods are being used, by and large most RANS applica-
tions today are still spatially second order, particularly for industrial applications.
Many researchers solve the turbulence equations uncoupled from the mean-flow
equations, and often the models are solved at second-order (or lower) spatial
accuracy regardless of the mean flow scheme. Although conventional wisdom
says that this method is reasonable, there may be more stringent requirements if
overall higher-order accuracy is desired.
RANS is inherently limited by the averaging procedure that defines it. Because
RANS is designed to model turbulence across the entire wave number energy
spectrum (unlike LES, which models turbulence only at the smallest scales above
some cut-off wave number), RANS is in effect trying to model the effects of all
turbulent eddies, even the large ones, in an average sense. As many researchers are
learning, there are some situations (like separated shear layers) where this approxi-
mation is poor, and the averaged results do not agree well with what can be
obtained by allowing the larger eddies to develop and interact on their own. Many
researchers refer to this characteristic of RANS as being overly dissipative, but
really it is simply the models trying to obtain an average through the action of
their eddy viscosity, precisely as they have been developed to do. It therefore
seems that RANS turbulence models in theory could be “fixed” to yield time-
averaged results in better agreement with measurements in those regions where
they currently fail.

D. Other Computational Fluid Dynamics Models and Methods


1. Blended RANS-LES Modeling
A new series of techniques to emerge over the last decade is based on the
blending of the RANS and LES equations. These were developed based on the
MODELING AND SIMULATION 201

idea that LES is still too expensive, particularly in the near-wall region for attached
boundary layers, and RANS typically performs poorly in large separated regions.
The idea is to blend the two methods, making use of the advantages of both.
RANS is used in boundary layers near walls, and LES is used outside this and in
separated regions. Because the filtered LES equations and the RANS equations
are identical up to the point of the subgrid scale (or turbulence) modeling, it is a
relatively simple matter to develop new “blended” models at this level.
One very popular method has been the detached eddy simulation (DES) tech-
nique developed by Spalart et al. (1997). The model is based on the RANS
Spalart–Allmaras model (Spalart and Allmaras, 1994), with the exception that the
distance to the nearest wall in the model is replaced by

dˆ = min(d , CDES D ) (58)


__ __
where D = max(Dx, Dy, Dz) and CDES is a constant. When d > CDESD, the Spalart–
Allmaras model (when its production and destruction__ terms balance) assumes a
form similar to the Smagorinsky model, nt µ D2S.
Other blended RANS-LES models have also been developed (Batten et al.,
2002; Baurle et al., 2001; Bush and Mani, 2001; Girimaji and Lavin, 2006; Menter
et al., 2003; Spalart et al., 2006; Zhang et al., 2000), but in the interest of space they
will not be discussed here. These types of methods have been overall very success-
ful for applications with massive separation, but they tend to be problematic for
applications to attached or only mildly separated flows. There have been several
applications of blended RANS-LES models to flow control problems to date (Arad
et al., 2006; Arunajatesan et al., 2002; Gross and Fasel, 2006b; Hiller and Seitz,
2006; Israel et al., 2004; Krishnan et al., 2006; Paterson and Baker, 2006; Saric
et al., 2006; Spalart et al., 2003). These methods have generally shown mixed suc-
cess. For example, in Spalart et al. (2003), the DES model applied to a stalled airfoil
captured the flowfield well when flow control was off, but failed to capture the
effects of a synthetic jet with sufficient accuracy. A brief discussion of a few of the
other RANS-LES applications is given in Sec. III on CFD Validation.
It is important to note that use of a blended RANS-LES model implies that the
computation must be done in 3-D and time-accurately. This is because use of LES
denotes that the simulation is resolving the large eddy structures, which are
unsteady 3-D features. Thus, blended RANS-LES simulations are still more
expensive than most routine 2-D RANS runs by a very large factor, and are con-
sequently not typically employed for routine industrial applications.
One problem with blended RANS-LES models that switch between the two
methods based on some grid-related measure is that the solution can be highly
dependent on grid quality, and often requires strict rules for grid generation in
order to obtain a consistent solution. For example, it is possible to cause “grid-
induced” separation if the grid aspect ratio gets too low in the boundary layer, as
discussed in Menter et al. (2003). Also, if the switch occurs too low in the bound-
ary layer, a noticeable mismatch in the log-law profile can be produced (Piomelli
et al., 2002). Some of the more recent blended models have been designed in an
attempt to overcome these problems. A great deal of research in blended RANS-
LES models is ongoing.
202 C. L. RUMSEY AND R. C. SWANSON

2. Reduced-Order Modeling
So-called reduced-order modeling has been shown to be useful for flow control
applications with actuators, as described in Yamaleev and Carpenter (2006). In
this model, the flow inside an actuator plenum is simulated by using the time-
dependent compressible quasi-one-dimensional Euler equations

∂Q ∂F
+ +T=0 (59)
∂t ∂z

where Q represents the vector of conserved variables, Q = (A/J)[r, rv, rE]T, J is


the Jacobian of the coordinate transformation, A is the cross-sectional area of the
quasi-one-dimensional representation of the actuator, and

È z t + z y rv ˘
AÍ ˙
F = Í z t rv + z y ( rv + p) ˙
2
(60)
JÍ ˙
Îz t r E + z yv( r E + p)˚

È 0 ˘
Í ˙
1 ∂A
T = - Íp ˙ (61)
J Í ∂y ˙
Í ˙
Î 0 ˚

The diaphragm oscillations are specified, and a moving mesh technique is used to
solve the equations. This reduced-order model inside the plenum can then be
combined with the RANS equations near and outside the jet exit. Use of this
model can simplify computations when the plenum is geometrically very com-
plex, and RANS simulations of the interior are impractical.
A few other low-order models are mentioned here, but are not described in
detail. Gallas et al. (2003) used the method of lumped element modeling for
piezoelectric-driven actuators. This model makes use of an equivalent electrical
circuit to obtain the volume flow rate through the orifice. Yamaleev and Carpenter
noted that lumped element modeling neglects compressibility effects associated
with large pressure oscillations in the plenum. Ito and Ravindran (1998), Park and
Lee (1998), Ravindran (2000), and Gross and Fasel (2007d) all developed reduced-
order models for the incompressible Navier–Stokes equations. The goal of these
models was to capture the essential flow control physics in the vicinity of the
design operating point by using either a reduced-basis or a reduced set of empiri-
cal eigenfunctions, derived from the experimental or numerical data of a system.
The latter two references make use of proper orthogonal decomposition (POD), a
technique which is optimal in the sense that it maximizes the amount of energy
captured for a given number of modes.

3. Immersed Boundary Methods


In this subsection we turn away from topics related to general equation meth-
odologies and briefly examine the specific technique of immersed boundary
MODELING AND SIMULATION 203

methods in some detail. Such methods could possibly have a significant impact on
computational capability for flow control, since they employ Cartesian grids and
thus relax the grid generation requirement. They could be applied to any of the
sets of governing equations discussed so far, including DNS, LES, and RANS.
One of the demanding requirements in CFD is generating an appropriate grid (i.e.,
adequate resolution) for flow over complex geometries. The degree of difficulty of
this requirement can be augmented considerably for flow control problems. For
example, generating a suitable grid when several actuators for different types of
flow control are being applied or an array of flow control devices is being employed
can be especially challenging. A possible method for overcoming this particular
challenge may be the immersed boundary (IB) technique.
Over the last decade considerable effort has been expended in the development
of the IB method, which was introduced by Peskin (1972) to study blood flow
passing through the mitral valve of the heart. With this method a Cartesian grid is
used to cover the entire domain of interest, including the geometry being consid-
ered. In general, the grid does not conform to the boundaries of the domain, and
the boundary conditions cannot be applied directly. An alternative approach is
required to impose the boundary conditions such that their effect on the flow is
accurately represented. The choice of an appropriate approach for imposing the
boundary conditions is crucial to the success of the IB method. There are several
salient advantages of the IB technique. One of the primary advantages is the use
of a Cartesian grid, which is much simpler to generate and does not have skew-
ness. The grid distortion that often occurs in body-conforming grids for complex
geometries can adversely effect both accuracy and the convergence behavior of
the flow solver. When the grid is Cartesian, the implementation of different types
of discretization such as higher-order approximations is also simpler (e.g., no
curvilinear coordinate transformations). Another principal benefit of the IB
method is the elimination of grid deformations and remeshing strategies for mov-
ing bodies. While these advantages make the IB method quite attractive, there are
certain disadvantages that must also be considered. The disadvantages primarily
center around the accuracy of the IB treatment and the requirement for high
Reynolds number flows. Conservation properties in the vicinity of the IB may also
need special attention. These issues will be addressed in subsequent discussion.
For the IB treatment there are two principal techniques for imposing the bound-
ary conditions: 1) appending an additional term, which is called a forcing function,
to the governing equations, and 2) modifying the discrete numerical scheme near
the boundary. The first technique employs a continuous forcing function, while
the second applies a discrete forcing function. In the subsections to follow we
succinctly describe these techniques and provide some discussion of the advan-
tages and disadvantages of each one. Initially, we restrict our attention to second-
order spatial discretizations, but there is some discussion on recent efforts to
implement higher-order discretizations. We conclude this section by presenting
some examples of applications of the IB method to flow control problems.
a. Continuous Forcing Approach. Consider a body immersed in a domain
W containing the subdomains Wf and Wb, and let the body occupy the domain Wb
with boundary Gb. Let the governing flow equations be written as

LU = 0 in W f (62)
204 C. L. RUMSEY AND R. C. SWANSON

U = UG on G b (63)

where L is the differential operator and U is the solution vector. If the domain
W is covered with a Cartesian grid, as shown in Fig. 4, and Eq. (62) is discretized
on that domain, the boundary conditions of Eq. (63) cannot in general be applied
directly. In order to provide the effect of the boundary, a forcing function (source
term) fb is introduced into Eq. (62). Then, Eq. (62) is rewritten as

LU = fb (64)

which applies to the entire domain (Wf + Wb). On a Cartesian mesh covering the
complete domain, we discretize Eq. (64). Introducing the forcing function into the
governing equations before discretization has the advantage that the function does
not depend on the type of the discretization.
As an example of the continuous forcing approach we consider a flow with
elastic boundaries, following the discussion given by Mittal and Iaccarino (2005)
related to the IB methods proposed by Peskin (1972, 1981). Suppose that the
domain is covered with a fixed Cartesian mesh and the immersed boundary is
represented by a set of elastic fibers which are tracked with a Lagrangian approach
(i.e., the fibers move with the local flow velocity). Let Xk be the coordinate vector
of the kth Lagrangian point on a fiber. The time rate of change of the location of a
point on the fiber is equal to the fluid velocity at that point, giving the equation
∂Xk
____ = u(Xk, t) (65)
∂t

Γb

Ωb

Ωf

Fig. 4 Boundary of geometry Gb immersed in Cartesian grid.


MODELING AND SIMULATION 205

The stress can be determined by a constitutive relation such as Hooke’s law.


The forcing function that represents the effect of the fiber stress on the fluid is
given by

fm (x, t ) = Â Fk (t ) d (| x - X k |) (66)
k

where d is the Dirac delta function. This forcing function is added to the momentum
equations. Since the immersed boundary does not generally coincide with the
Cartesian grid points, the forcing function cannot be transferred directly to the
Cartesian grid points. Thus, the sharp d function must be replaced by a smoother
distribution function. A replacement distribution function is designated d, and it
spreads the forcing over the group of cells surrounding each Lagrangian point. Then
the forcing at each grid point xi, j due to the elastic forces in the fibers is given by

fm (x i , j , t ) = Â Fk (t ) d (| x i , j - X k |) (67)
k

The fiber velocity in Eq. (65) can also be determined with the same distribution
function. The success of this method depends upon the choice for the distribution
function d. This method has been successfully applied to a variety of problems
involving elastic boundaries.
For rigid boundaries, a continuous forcing function could be determined with
the approach just described by considering the body to be elastic but extremely
stiff. One alternative approach is to consider the immersed boundary attached to
an equilibrium position by a spring with a restoring force given by

Fk = -k(Xk - Xek(t)) (68)

where k denotes the spring constant and Xek(t) is the equilibrium position of the kth
Lagrangian point. To approximate the boundary as rigid requires a large spring
constant, which makes the system of equations stiff and more difficult to solve.
As indicated by Mittal and Iaccarino (2005), several other methods have been
considered to represent the influence of a rigid IB on the fluid using the continu-
ous forcing function, including the feedback control method of Goldstein et al.
(1993), which has the desirable property of adjusting the forcing with the evolu-
tion of the solution. These methods generally produce stiff numerical systems. In
addition, they do not allow a sharp representation of the IB; and thus they are not
well suited for turbulent flow simulations.
b. Discrete Forcing Approach. The second technique for imposing bound-
ary conditions in the IB method is the discrete forcing approach. With this approach
Eq. (64) is first discretized on a Cartesian grid covering the entire domain, includ-
ing the body. In the cells near the immersed boundary the discretization is adjusted
to account for the effect of the boundary conditions, producing the modified
system of equations

L¢U = r (69)
206 C. L. RUMSEY AND R. C. SWANSON

where L¢ is the modified discrete operator, and r is the vector of known terms that
depend on the boundary conditions on the immersed boundary Gb. The discrete
system of Eq. (69) can be rewritten as

LU = f b¢ (70)
where f b¢ = r + LU - L¢U. As revealed in subsequent discussion, f b¢ in effect
emerges from the discretization near the boundary. That is, with the discrete
approach the forcing function depends on the discretization. We now consider
three methods of applying the discrete forcing approach: 1) Mohd–Yusof method,
2) the ghost-cell finite-difference method, and 3) the cut-cell finite-volume
method.
With the continuous forcing method there is generally a severe time-step limita-
tion due to the stiffness of the flow equations that arises because of parametric
values [e.g., the spring constant in Eq. (68)] required to represent rigid boundar-
ies. Mohd-Yusof (1997) proposed a discrete forcing method that would eliminate
this severe time-step restriction and prevent possible spurious oscillations when
the forcing function is not distributed over several grid points. In this method a
forcing function can be defined such that the boundary conditions are imposed at
the immersed boundary. More specifically, consider a discretized form of the
Navier–Stokes equations written as

ruin+1 - ruin
= RHSi + fi , i = 1, 3 (71)
Dt

where Dt denotes the time step, and RHSi represents the convective, diffusive
pressure gradient, and source terms of a momentum equation. Assuming that the
grid point being considered is coincident with the immersed boundary and that the
Dirichlet boundary condition (rui) = (rui)s is imposed, then the forcing function
in Eq. (71) to ensure direct enforcement of the boundary conditions is given by

( rui )s - ruin
fi = - RHSi (72)
Dt

The application of Eqs. (71) and (72) means that there is actually no need to
compute the forcing function. However, the grid points do not in general coincide
with the IB, and we need an interpolation procedure for calculating the unknowns
adjacent to the IB so that the boundary conditions are enforced.
Mohd-Yusof has investigated several ways to determine the velocities at the
fluid solution points adjacent to the body and at the interior points of the body.
One way is to use a linear interpolation that gives the specified velocity at the IB.
This type of interpolation procedure can also be used to obtain the velocities at the
interior points. For example, at an interior point adjacent to the boundary one can
reverse the velocity so that a linear extrapolation gives the appropriate velocity at
the IB. For arbitrary curved surface boundaries another type of interpolation (e.g.,
2-D linear, quadratic) is necessary. If the flow variables are not varying smoothly
near the boundary, Iaccarino and Verizicco (2003) indicate that the inverse
weighted method of Franke (1982) may be appropriate, since this method has the
property of preserving local maxima and producing smooth reconstruction.
MODELING AND SIMULATION 207

Another important element in the implementation of Mohd-Yusof (1997) is


related to the use of a Fourier pseudospectral method. Even though a reasonably
smooth variation in velocity is achieved in the vicinity of the IB, the rather sharp
variation in velocity can produce spurious oscillations. To alleviate this problem a
negative velocity is created at the first solution point inside the body by extending
the interpolation just discussed. Moreover, there is a certain smoothing that occurs
by solving the associated interior flow. It should be emphasized that the desired
flow solution outside the body is essentially independent of the internal conditions.
In the second discrete forcing method the boundary conditions on the immersed
boundary are imposed through the use of ghost cells, which are defined on the
inside of the body such that they have at least one neighbor exterior to the body,
as shown in Fig. 5. Interpolation schemes can be defined that include the bound-
ary conditions on the IB. One simple example is bilinear (trilinear in 3-D) inter-
polation; a generic flow variable f can be expressed as

f = C1xy + C2x + C3y + C4 (73)


where the coefficients C1 through C4 can be determined by evaluating f at the fluid
nodes F1, F2, and F3 (see Fig. 5), and at the boundary point B2 between the points
P1 and P2. The point B2 is the normal intercept with the line that passes through the
ghost point G. Note that P1 and P2 are the boundary intercepts of the constant y

Fig. 5 Ghost cell method for imposing the boundary conditions at the immersed
boundary; the fluid points F1 and F2, ghost point G, and the boundary point B2 (or B1)
are used to determine the interpolation scheme.
208 C. L. RUMSEY AND R. C. SWANSON

and x lines passing through the ghost point. The midpoint point B1 between P1 and
P2 can be used instead of B2. For laminar flows, and turbulent flows with a solution
point inside the viscous sublayer, a less accurate linear interpolation (i.e., C1 = 0)
is acceptable.
For any interpolation scheme the value of the ghost-cell unknown, designated
fG, can be expressed as

Âw f i i = fG (74)

with the summation extending over all the points in the computational stencil,
which includes one or more boundary points. The coefficients wi depend upon
geometric data (e.g., inverse distance weighting which depends upon the distance
between fi and fG, see Iaccarino and Verzicco, 2003). Equation (74), representing
Eq. (70) (the modified equation), can be solved simultaneously with the discretized
flow equations for the unknowns. This method has been successful for both
incompressible and compressible viscous flows with Reynolds numbers up to
O(105) (e.g., Ghias et al., 2004; Tseng and Ferziger, 2003).
With the ghost-cell method for discrete forcing there is no direct attempt to
guarantee satisfaction of the conservation laws for the cells near the immersed
boundary. In the third discrete forcing method, by combining a boundary cut-cell
method with a finite-volume formulation, the forcing function can be determined
such that a discrete conservation of mass, momentum, and energy is enforced. The
first steps in applying this method are to identify the cells cut by the immersed
boundary and to determine the intersection of the sides of the cut cells with the IB.
The finite-volume approach requires the reshaping of the cut cells so that the IB
coincides with a cell face. If the center of the cell is in the fluid, the portion of the
cell inside the body is discarded. Otherwise, the cut cell can be merged with a
neighboring cell. Upon completing this process the resulting surface control vol-
umes have a trapezoidal shape, as shown in Fig. 6a.
In a finite-volume formulation we need to approximate the flux integrals for
mass, momentum, and energy fluxes at each face of a cell. This can be done by
devising an appropriate interpolating polynomial for a flow variable f that is valid
in the region where the flux is required, and then evaluating the flux. Ye et al.
(1999) proposed a 2-D polynomial interpolating function for this purpose. As an
example of this approach, consider the flux fsw indicated in Fig. 6. This flux can be
approximated by using the six-point interpolation stencil

f = C1xy2 + C2y2 + C3xy + C4x + C5y + C6 (75)


which is linear in x and quadratic in y. The coefficients C1 to C6 can be determined
by evaluating f at the six points marked by squares in Fig. 6b. As pointed out by
Mittal and Iaccarino (2005), Eq. (75) represents the most compact function that
allows the evaluation of f or its derivative at the sw face with at least a second-
order accuracy. The other fluxes can be computed in a similar manner.
There are several benefits derived from the application of the cut-cell finite-
volume method. The method not only ensures the satisfaction of the conservation
properties but also eliminates the requirement to compute information inside the
solid body. In addition, the resulting discrete forcing allows a sharp representation
MODELING AND SIMULATION 209

Fig. 6 Cut cell method for imposing the boundary conditions at the immersed
boundary (IB); a) reshaped cell adjacent to IB and associated fluxes; b) interpolation
stencil based on 6 points for determining flux at the southwest face ( fsw).
210 C. L. RUMSEY AND R. C. SWANSON

of the IB effects (i.e., no distribution of the forcing) on the flow solution, which is
necessary for high Reynolds number flows. The cut-cell method has been success-
fully applied to a variety of complex 2-D problems, for example, flow induced
vibrations, flapping airfoils, and multiple objects in free fall through a fluid.
However, extension of the method to 3-D is not straight forward, since complex
polyhedral cells are created, making the discretization of the full Navier–Stokes
equations much more difficult.
The discrete forcing approach provides the advantage of direct control over the
accuracy, stability, and conservation properties of the flow solver. For the ghost-
cell and cut-cell methods there is no requirement to compute the flow throughout
the interior of the body, resulting in a reduced computational effort. Thus, there is
some relaxation of the disadvantage of IB methods relative to body-conforming
methods when computing high Reynolds number flows. With the discrete forcing
function approach, inclusion of boundary motion can introduce some additional
complications. However, rather than recomputing the body location for each time
step, one can attach the computational grid to the body and account for the motion
of the fluid relative to the moving Cartesian grid (see Cho et al., 2007). For more
detailed discussion of the various elements of the IB method see reviews by
Iaccarino and Verzicco (2003) and by Mittal and Iaccarino (2005).
c. Example Applications. The IB technique has been applied primarily to
incompressible flow problems. For these incompressible applications second-order
and fourth-order spatial accuracy as well as spectral accuracy have been considered
(e.g., Fadlun et al., 2000; Linnick and Fasel, 2005; Mohd-Yusof, 1997). However,
in the last several years the IB method has been used in some compressible flow
computations, such as those performed by Von Terzi et al. (2001), De Palma et al.
(2006), and Cho et al. (2007). Both Von Terzi and Cho have used higher-order
spatial differencing (fourth-order compact and fifth-order Weighted Essentially
Non-Oscillatory (WENO), respectively) and higher-order temporal approxima-
tions. Von Terzi has demonstrated that the simulation of Tollmein–Schlicting
waves in a flat plate boundary layer is a good test case for evaluating the accuracy
of an IB method in the vicinity of the boundary. In addition to these applications
the IB approach has also been used by Kellogg (2000) and You et al. (2006a) to
solve some flow control problems.
Kellogg (2000) obtained LES solutions to several control problems using the
IB method. In these applications he solved the filtered incompressible Navier–
Stokes equations, computing the subgrid scale stress tensor with the Smagorinsky
model. Using a staggered grid approach he employed a Fourier spectral method
for discretization in two coordinate directions (imposing periodicity in those
directions) and a second-order finite difference approximation in the wall-normal
direction. For temporal discretization the wall-normal convection and diffusion
terms were treated implicitly with Crank–Nicolson, and other terms were advanced
in time with a third-order Runge–Kutta method. The system of discrete equations
was solved by a fractional step method, a two-step procedure satisfying the
divergence-free condition after the second step. One of the problems Kellogg
considered, which demonstrated the flexibility of the IB method, involved applying
an opposition control scheme to cancel the coherent turbulent structures in a
channel flow. This opposition control scheme used the motion of a wall to produce
a normal flow velocity to cancel an opposing velocity in the interior of the flow.
MODELING AND SIMULATION 211

Kellogg also considered a control problem with micro-electromechanical system


(MEMS) actuators for a closed loop, active control of turbulent flow. He per-
formed simulations for drag reduction that modeled both single and multiple
MEMS-like actuators, again demonstrating the capability of the IB method.
You et al. (2006a) performed LES computations using the IB method for
incompressible turbulent flow with an embedded streamwise vortex pair produced
by two wall-mounted, half-delta wings (i.e., vortex generators). The spatial dis-
cretization used was second-order in each coordinate direction, and the time
advancement procedure was similar to that used by Kellogg. With the IB method,
the complex flow features observed in experiments were captured. In addition, the
computed mean flow velocity and Reynolds stress profiles compared favorably
with experimental data.

III. Computational Fluid Dynamics Validation


Clearly, one of the most important aspects of modeling and simulation for any
field, including active flow control, is CFD validation. Models and methods must
be compared against experiment, theory (if possible), and each other in order to
validate them. This is the process that advances the state of the art. Unfortunately,
most CFD validation efforts tend to be isolated from one another. The greatest
strides forward are often made when many researchers validate and compare
results and methods on the same problem. When it is possible for experimentalists
and CFD researchers to work together to define both experiments and simulations,
the benefits can be even greater. This section summarizes two recent workshops
specifically for active flow control (Jones and Joslin, 2005; Rumsey et al., 2006,
2007b), and also mentions an ERCOFTAC-sponsored series of workshops on
Refined Turbulence Modelling. There have been other flow control-related meet-
ings and symposia,* but the workshops to be described in detail are the only recent
ones the authors are aware of for which workshop participants computed the same
or similar flow control test cases, and for which thorough documentation is easily
accessible.

A. Summary of Circulation Control Workshop


With circulation control (CC), a tangential wall jet is used primarily for the
purpose of enhancing lift of an aerodynamic surface. This type of active flow
control can provide substantial benefits for real-world applications, as demonstrated
for the V-22 tiltrotor aircraft configuration (Jacot and Mabe, 2000). Organizers of

*Other recent flow-control meetings include, for example the Workshop on Flow Control:

Fundamentals and Practice in Corsica, 1996; IUTAM Symposium on Mechanics of Passive and Active
Flow Control in Gottingen, 1998; 4th SIG 33 Workshop on Flow Control in Abisko, 2001; 9th
ERCOFTAC/IAHR/COST Workshop on Refined Turbulence Modelling in Darmstadt, 2001; AFOSR
Workshop on Plasma Actuators for Subsonic Applications at Eglin AFB, 2004; Active Control of
Aircraft Noise—Concept to Reality in Stockholm, 2005; IUTAM Symposium on Flow Control and
MEMS in London, 2006; First Berlin Conference on Active Flow Control, 2006; and any of the AIAA
series of Flow Control meetings held biennially (to date) in 2002, 2004, and 2006.
212 C. L. RUMSEY AND R. C. SWANSON

LIVE GRAPH
Click here to view

Fig. 7 View of typical circulation control airfoil near trailing edge.

the 2004 Circulation Control Workshop (Jones and Joslin, 2005) held in Hampton
Virginia asked CFD participants to compute flow over the NCCR 1510-7607 N
airfoil, with blowing over its circular (Coanda) trailing edge. This CC experiment
was conducted in 1977 by Abramson (1977). Some participants computed flow
over a similar configuration, the 103RE(103XW) airfoil, tested by Abramson and
Rogers (1983). A graphic showing the trailing edge of the airfoil configuration is
shown in Fig. 7. The wall jet emanating from the plenum “sticks” to the trailing
edge surface due to the Coanda effect, causing delayed flow separation and thus
increasing circulation and producing higher lift. For the blown conditions, all
participants at the workshop used RANS and a variety of turbulence models,
including one-equation linear, two-equation nonlinear, and full Reynolds stress.
One surprising result to come out of the workshop was the inconsistency in the
CFD results. In particular, many RANS computations could—for some blowing
conditions—obtain unphysical results, where the jet wrapped around the lower
airfoil surface. An example is shown in Fig. 8. Furthermore, even when different
computer codes used (ostensibly) the same turbulence model, significant differ-
ences were seen between the reported results.
Many recent papers have been published on CC airfoil flows (Baker and
Paterson, 2006; Chang et al., 2005, 2006; Fasel et al., 2006; McGowan and
Gopalarathnam, 2006; McGowan et al., 2006; Paterson and Baker, 2004, 2006;
Slomski et al., 2002; Swanson et al., 2005, 2006; Swanson and Rumsey, 2006
Zacharos and Kontis, 2004). The bottom line seems to be that a) for RANS, some
turbulence models—such as Menter’s k-w SST (Menter, 1994), SARC (Spalart
and Shur, 1997), k-z (Warren and Hassan, 1998), explicit algebraic stress (Rumsey
and Gatski, 2001), and full Reynolds stress models (Launder et al., 1975)—can do
well for certain conditions, but it depends on the case and there is a tendency for
MODELING AND SIMULATION 213

LIVE GRAPH
Click here to view

Fig. 8 Example showing Coanda jet wrapping around the airfoil nonphysically far.

the solutions to degrade compared with experiment as the blowing increases; and
b) these CC flows tend to be very sensitive to numerical parameters. For example,
Swanson and co-worker demonstrated high sensitivity in the SST model to how
the production term was computed (Swanson et al., 2005) as well as to grid density
(Swanson and Rumsey, 2006). Other types of computations, including DES
(Paterson and Baker, 2006) and LES (Slomski et al., 2006) have been rather iso-
lated and too preliminary to draw any firm conclusions about their potential for
this class of CC flows.
It also became evident at the workshop that more CC experiments are needed
for validation. This workshop used data from experiments conducted nearly 30
years ago! It was further recognized that this type of experiment is difficult to
perform, particularly when trying to maintain two-dimensionality as blowing
increases. At this point we will inject a comment concerning the jet momentum
coefficient for CFD validation. In experiments, the jet momentum coefficient is
often the only information given regarding a very crucial aspect of the CFD simu-
lation: the jet boundary condition. It is defined as

 j
mV
Cm = (76)
1
2 r•V•2 L

where the jet velocity Vj is typically obtained from conditions inside the plenum
combined with isentropic flow relations. Unfortunately, this methodology intro-
duces a degree of uncertainty into the CFD simulations. It is far more useful to the
CFD community to have measured values of the flowfield at the jet exit plane,
including velocity (and turbulence properties if possible).
214 C. L. RUMSEY AND R. C. SWANSON

B. Summary of CFDVAL2004 Workshop


The Langley Research Center Workshop on CFD Validation of Synthetic Jets
and Turbulent Separation Control (also known as CFDVAL2004) was held in
Williamsburg, Virginia in 2004 (Rumsey et al., 2006, 2007b). This workshop was
unique in that it brought together both experimentalists and CFD experts, and the
three flow control experiments were designed and performed specifically for the
workshop. Furthermore, the experimental data and workshop results were posted
to a public website,* which has subsequently encouraged a great deal of additional
research and published papers on the test cases. To date, there have been nearly 40
CFD papers published related to the CFDVAL2004 workshop. A partial list—of
only the journal articles published to date—is given here (Carpy and Manceau,
2006; Cui and Agarwal, 2006; Dandois et al., 2006a; Kotapati et al., 2007;
Krishnan et al., 2006; Morgan et al., 2006; Postl and Fasel, 2006; Rumsey, 2007a, b;
Saric et al., 2006; Vatsa and Turkel, 2006; Yamaleev and Carpenter, 2006; You
et al., 2006b). A complete list of all papers (some of which are precursors to the
journal articles) can be found on the website.
The three workshop cases were chosen to represent different aspects of flow
control physics: nominally 2-D synthetic jet into quiescent air, 3-D circular syn-
thetic jet into turbulent boundary-layer crossflow, and nominally 2-D flow-control
(both steady suction and oscillatory zero-net-mass-flow) for separation control on
a simple wall-mounted aerodynamic shape. It is important to note that experiments
are difficult to perform for these types of unsteady flowfields. In the experiments
here, an effort was made to take duplicate measurements using different techniques;
this duplication highlighted the uncertainties inherent in the measurements
(Greenblatt et al., 2006a, b; Naughton et al., 2006; Schaeffler and Jenkins, 2006;
Yao et al., 2006). A summary of the workshop results can be found in Rumsey et al.
(2006). Here we briefly recap the main conclusions, then summarize results of
additional CFD research that has occurred subsequent to the workshop.
Case 1, 2-D synthetic jet into quiescent air, was a difficult experiment to simu-
late. The flowfield was probably partially laminar or transitional, so it was unclear
how best to simulate it. Workshop participants used RANS, laminar Navier–Stokes,
blended RANS-LES, LES, and a reduced-order model. End effects probably
caused significant three-dimensionality far away from where the jet emanated
from the wall, but most participants computed the flow in 2-D. The piezoelectric
driver was a difficult device to simulate, so most computations made approxima-
tions inside the plenum or simply applied jet boundary conditions directly on the
wall from which the jet emanated. As a result, the simulations did not even start
off with the same conditions as the experiment at the jet exit: deviations from
periodicity were for the most part not simulated. For RANS, it was therefore dif-
ficult to judge the capabilities of the turbulence models. However, Carpy and
Manceau (2006), who used extracted PIV data near the slot exit as surface bound-
ary conditions, later noted that full Reynolds stress models offer improvements
over linear models when representing turbulence dynamics for this case, because
they can capture the presence of a region of negative production that occurs during
the deceleration phase. They also brought up the point that when applying surface

*http://cfdval2004.larc.nasa.gov [cited 9/2007].


MODELING AND SIMULATION 215

boundary conditions for the jet (i.e., not solving the flow in the cavity), prescrib-
ing an inlet value for e can be problematic.
It is important to note that additional measurements taken after the workshop,
given in Yao et al. (2006), were at somewhat different conditions from those used for
the workshop case. Some of the larger discrepancies exhibited by different measure-
ment techniques in the original data were mitigated in the later experiment.
Subsequent investigations performed with 2-D RANS by Vatsa and Carpenter
(2005), Vatsa and Turkel (2006), and Park et al. (2007) indicated generally
improved comparisons with experiment. The primary reason for the improvement
was likely the increased attention given to achieving similar velocity profiles to
experiment at the jet exit. All of these computations modeled a simplified plenum
with periodic transpiration applied on the bottom wall, but the imposed velocity
was curve-fitted to better replicate the experiment. In particular, Vatsa and co-
workers obtained nearly the precise temporal variation of the experimental signal
by curve-fitting the measured velocities at the slot exit with a fast Fourier trans-
form to reflect the proper mode shapes and to ensure zero net mass. As a result,
they were able to obtain flowfield results in very close agreement with experiment.
An example using the Spalart–Allmaras turbulence model is seen in a plot of
time-averaged vertical velocity along the jet centerline in Fig. 9.
Simulations using the 3-D incompressible Navier–Stokes equations have been
performed by Kotapati and Mittal (2005) and Kotapati et al. (2007). They mod-
eled an approximate plenum shape with periodic spanwise boundary conditions,
and claimed that their computations captured the transitional nature of the flow-
field. Xia and Qin (2006) performed 2-D laminar and 3-D DES simulations with
LIVE GRAPH
Click here to view

Fig. 9 Time-averaged vertical velocity along the jet centerline for CFDVAL2004
Case 1 (synthetic jet into quiescent flow), (from computations of Vatsa and Turkel,
2006).
216 C. L. RUMSEY AND R. C. SWANSON

periodic spanwise boundary conditions. Unlike most others, they attempted to


model the geometry of the piezoelectric driver (in a 2-D sense) with greater fidel-
ity and included a wave-like function on the velocity profile applied at the plenum
side wall where the driver was located in the experiment. However, resulting mean
velocity profiles at the jet exit were not as accurate compared with the technique of
Vatsa and Turkel (2006). These 3-D simulations indicated that 3-D Navier–Stokes
(essentially an under-resolved direct simulation) and blended RANS-LES meth-
ods are capable of simulating this type of flowfield. Unlike RANS, these methods
hold the promise of more accurately predicting turbulence effects, through directly
resolving the large eddy structures. It is not known whether accurate prediction of
these turbulence effects for this type of flowfield is required for adequate perfor-
mance predictions. In any case, when comparing against experiment, faithfully
mimicking the boundary conditions at the jet exit is certainly of primary impor-
tance for all methods: RANS, LES-type, and direct simulations alike.
Case 2, 3-D circular synthetic jet into turbulent crossflow boundary layer, was
the least computed of the three workshop cases, probably because it was necessarily
3-D. Most workshop participants used RANS, and one used LES. The experiment
exhibited a large cross-flow velocity (with peak the same order of magnitude as V•)
of unknown origin at the jet orifice exit, which was not modeled in any of the CFD
simulations. Qualitative agreement with experiment was reasonably good, but
quantitative comparisons showed significant variations. Different turbulence mod-
els were found to have less of an impact than different grids, codes, or other solu-
tion variants. Somewhat unexpectedly, LES and RANS solutions on similar-sized
grids yielded very similar results in mean-flow quantities. However, as described in
Dandois et al. (2006a), LES gave better turbulent stress predictions.
Other work appearing subsequent to the workshop (Biedron et al., 2005; Cui and
Agarwal, 2005; Iaccarino et al., 2004; Rumsey, 2004, 2007a) were mostly RANS
[although Cui and Agarwal (2005) also tried DES], and all used similar boundary
condition methodologies (periodic vertical velocity imposed on the bottom wall of
the plenum). Similar to the workshop, results among these papers seemed to vary
widely. However, the RANS methodology was certainly capable of obtaining very
good results for certain averaged quantities compared with experiment, as demon-
strated in Fig. 10. This figure shows the boundary layer perturbed in its phase-av-
erage by the passage of the synthetic jet structure. It indicates good agreement with
experiment and relative agreement between three different turbulence models and
two different grid sizes. It also demonstrates the degree of uncertainty inherent in
the experiment: there were fairly large differences between results obtained using
LDV and PIV at this particular location and phase.
Rumsey et al. (2007) computed a somewhat different circular jet into crossflow
(Milanovic et al., 2005) in addition to this one, and compared the two. Numerical
effects were explored, such as the effect of grid size, time-step, number of subitera-
tions, symmetry vs full plane, and effect of imposing jet boundary conditions on
the floor. For the latter, it was found that use of a top-hat sinusoidal boundary
condition at the orifice exit plane was an oversimplification that failed to capture
the complex nature of the flowfield near the orifice. A recommendation was made
to include at least some portion of the orifice in the computation. This conclusion
is in agreement with earlier work by Rizzetta et al. (1999), who found that account-
ing for the internal actuator geometry in a synthetic jet computation was important
because it affected the jet profiles at the exit.
MODELING AND SIMULATION 217

Fig. 10 Profiles of phase-averaged u-velocity on wall 1-diameter downstream of


orifice for CFDVAL2004 Case 2 (synthetic jet into crossflow), at phase = 120 deg: a)
comparison of CFD with experiment (LDV) and b) comparison of LDV with PIV
measurements.

Case 3 was flow over a nominally 2-D wall-mounted hump, inspired by the
earlier experiments of Seifert and Pack (2002). The Reynolds number was near 1
million based on chord. On this model, the flow (with no control) separates near
65% chord, and reattaches downstream past the end of the hump. Either steady
suction control or oscillatory synthetic jet control applied near the separation
point can lessen the size of the separation bubble. After the CFDVAL2004 work-
shop, the experimental data from case 3 were included as part of the ERCOFTAC
on-line database (Classic Collection),* and the data were also included as test
cases in two subsequent workshops: the 11th ERCOFTAC/IAHR Workshop on
Refined Turbulence Modelling in Goteborg, Sweden, 2005, and the 12th
ERCOFTAC/IAHR Workshop on Refined Turbulence Modelling in Berlin,
Germany, 2006. Although not discussed here, results from these later workshops
were consistent with the following discussion.
At the CFDVAL2004 workshop, 13 contributors ran 56 separate cases. Methods
were mostly RANS, but there were also blended RANS-LES results and one
under-resolved DNS result. Two important conclusions were made at the time
regarding this case. First, it was found that the side plates used in the experiment
caused blockage that needs to be accounted for in any CFD simulation, in order to
obtain reasonable wall pressures over the attached portion of the hump. Second,
nearly all models and methods at the workshop consistently predicted reattach-
ment location to be too far downstream. This can be seen in the plot of reattach-
ment location for the steady suction case in Fig. 11. Even the under-resolved DNS
predicted a separation bubble that was too long. Grid refinement studies and use

*http://cfd.mace.manchester.ac.uk/ercoftac/ [cited 9/2007].


218 C. L. RUMSEY AND R. C. SWANSON

Fig. 11 Reattachment location of CFDVAL2004 Case 3 (hump model) workshop


results with steady suction.

of methods with higher-order spatial accuracy did not help. Inside the bubble
itself, most RANS computations predicted velocity profiles in reasonably good
agreement with experiment, but underpredicted turbulent shear stress in magni-
tude. DES results at the workshop (Krishnan et al., 2006) (although not shown in
the figure because the reattachment point was not reported at the time) also showed
reattachment too far downstream for the suction case. However, results were bet-
ter for the no-flow-control case, and it was surmised that DES may need to seed
upstream eddy content into the boundary layer for shallow separations. Sarik et al.
(2006) also noted problems with DES for the shallow separation suction case, and
said that this was possibly due to sensitivity of the method to grid design and the
fact that the boundary layer upstream of separation was thinner.
Since the time of the workshop, aside from a greater emphasis on the oscillatory
(synthetic jet) case, no definitive progress has been made for RANS models in the
sense that new results (Balakumar, 2005; Bettini and Cravero, 2007; Capizzano
et al., 2005; Rumsey, 2007a) have mostly been consistent with results from the
workshop. But a good deal of work has been done in the area of blended RANS-
LES, LES, and (under-resolved) DNS. For the latter, Postl and Fasel (2006) doubled
the spanwise domain extent (from Dz/c = 0.071 used at the workshop to Dz/c =
0.142) and saw an improvement in their results, including reattachment length,
compared to experiment. Furthermore, unlike RANS methods, their results overall
did an excellent job of predicting turbulence levels in the separated region. Although
even at 200 million gridpoints the DNS simulation was still under-resolved at
MODELING AND SIMULATION 219
LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view

Fig. 12 Streamlines for CFDVAL2004 Case 3 (hump model) with steady suction:
a) recent LES computations of You et al. (2006b) and b) experiment (Greenblatt et al.,
2006a).

this Reynolds number, these computations demonstrated that this method can be a
useful tool when looking for insight into the flow physics.
Other blended RANS-LES and LES papers have been published (Hiller and
Seitz, 2006; Israel et al., 2004; Morgan et al., 2005a, b; Saric et al., 2006; You et al.,
2006b) for the hump model. Although the various results differed from each other
to some degree because of different subgrid models (or lack thereof for implicit
LES), different grid resolutions, and other factors, by and large these computations
demonstrated the capability for using LES methodology in the separated region to
improve predictions. In particular, by resolving many of the eddies in the separated
shear layer region and capturing the dynamics of the large-scale motion, these
methods produced increased levels of turbulence (and hence earlier reattachment),
in better agreement with experiment than RANS models. An example of a recent
comparison (You et al., 2006b) using LES is shown for the case with steady suction
in Fig. 12. The reattachment position in the computation is very close to the position
seen in the experiment. Although more costly to run than RANS, these LES methods
have become more affordable on today’s computers, and they offer the prospect of
obtaining a better understanding of the dynamics inherent in flow control problems
involving separation. It remains to be seen whether their use will lead to improved
RANS turbulence models for this class of flows.

IV. Exploration of Advanced Control Strategies


Using Computational Fluid Dynamics
In this section we consider several actual and possible active flow control tech-
niques and discuss numerical methods that have been used in simulations applying
these methods. Initially, we discuss models for the body force produced by plasma
actuators, since they are frequently used in computations to represent the effects of
the actuators on the flow. Then several examples of applications with plasma actua-
tors are given. In Sec. B we briefly discuss computational efforts to understand the
220 C. L. RUMSEY AND R. C. SWANSON

underlying mechanisms allowing polymers to produce drag reduction. This


understanding could provide the necessary insight to develop active flow control
techniques for drag reduction. The last three sections concern the use of CFD in
determining optimization of synthetic jet parameters and also possible feedback
control systems, including the application of neural networks.
This section is not intended to be an exhaustive representation of advanced
control strategies, but rather to provide the reader with a small sampling of
some interesting and promising emerging capabilities, with a primary focus on
the CFD methodologies behind them. Additional ideas and CFD methods—such
as applications to compliant coatings, bubble injection, electromagnetic control,
and wall oscillations—can be found in the literature (see, for example, Joslin
et al., 2005).

A. Plasma Actuators
One active flow control technique that is currently being investigated in a
variety of separation control applications involves dielectric barrier discharge
actuators (or plasma actuators) (see, for example, Roth et al., 2000). Plasma actua-
tors have two electrodes separated by a dielectric material. A high ac voltage is
applied to the electrodes causing the surrounding air to be ionized. The electric
field produced by the electrodes acts on the ionized air (plasma) to create a body
force vector which can induce steady and unsteady velocity components. The
desired type of body force acting on the fluid can be achieved by appropriate
design of the electrode geometry. For example, different electrode arrangements
can produce a jet in a specific direction, streamwise vortices, or spanwise vortices.
The body force created by a plasma actuator can be expressed as a function of
the applied voltage, and then included in the summation of forces in the Navier–
Stokes equations. Here we briefly discuss two types of body force models that
have been used when solving the Navier–Stokes equations. Detailed discussions
on these types of models are given by Orlov and Corke (2005), Shyy et al. (2002),
and Suzen et al. (2005). One type of force model is derived by considering
Maxwell’s equations and making certain simplifying assumptions. If we neglect
magnetic forces, the electrohydrodynamic force can be written as

fB = rc E (77)
where fB is the body force per unit volume, rc is the net charge density, and E is the
electric field vector. If the time variation of the magnetic field vector is also neglected,
one of Maxwell’s equations becomes — × E = 0. Since the electric field is irrota-
tional, it can be represented as the gradient of a potential function; and thus

E = -—F (78)
Using Maxwell’s equation for the electric induction vector, which is equal to
the product of E and the dielectric coefficient e (permittivity), and Eq. (78),
we obtain

rc
— ◊ (e—F ) = - (79)
e0
MODELING AND SIMULATION 221

where e0 is the permittivity of free space. The charge density rc can be expressed
as [Enloe et al. (2004)]

e0
rc = - F (80)
ld2

where ld is the Debye length, which is the characteristic length for electrostatic
shielding in a plasma. Substituting for charge density in Eq. (79)

F
— ◊ (e r —F ) = (81)
ld2

where er is the relative permittivity of the medium. Once Eq. (81) is solved for the
electric potential, then Eqs. (78) and (80) give the electric field and the charge
density, respectively. Then, the body force is computed with Eq. (77). We refer to
this model for the body force as Model 1. The body force needs to be computed
only once. For unsteady plasma actuators, the force will be activated according to
a duty cycle (i.e., frequency of application).
Another body force model is the one introduced by Shyy et al. (2002). This
model does not require solving for the electric potential to obtain the electric field.
Instead, the spatial variation of the electric field is determined by

Va
| E | = E0 - k1 x - k2 y, E0 = (82)
de

where Va is the applied voltage and de is the separation distance of the two elec-
trodes. The constants k1 and k2 are determined by the condition that the electric
field strength is the breakdown value at the interface boundary of the plasma and
fluid. The body force is expressed as

f b = q f Dtarc ec Ed b (83)

where qf is the frequency of the applied voltage, Dt is the time during which the
plasma discharge takes place, a is a factor to account for collision efficiency, ec is the
electron charge, and the parameter db is 1 when the electric field is above its critical
value and zero otherwise. The charge density is assumed to be uniform in the region
of plasma formation. This model for the body force is referenced as Model 2.
As examples of computations involving flow control with plasma actuators, we
consider flow over a turbine blade, wall-mounted hump model, and a wing sec-
tion. The primary emphasis in each of these cases is separation control. Suzen
et al. (2005) solved the RANS equations for flow over a low-pressure turbine
blade and applied separation control with a single plasma actuator. For this low-
speed laminar flow the Reynolds number based on the turbine cascade inlet veloc-
ity and axial chord length was 5 × 104. The actuator body force was computed
with Model 1, which requires solving for the electric potential. In the computa-
tions the actuator was located just upstream of the start of the separation region,
which according to the experiment occurs between the 0.7 and 0.95 chord
222 C. L. RUMSEY AND R. C. SWANSON

locations on the suction surface of the turbine blade. A sine wave voltage with a
frequency of 5 kHz and an amplitude of 5 kV was applied to the actuator. While
the separated flow was not eliminated, the streamwise separation extent was
reduced by more than 30%.
He and Corke (2007) computed solutions to the RANS equations for turbulent
flow over a wall-mounted hump model, which is one of the cases from the
CFDVAL2004 workshop (Rumsey et al., 2006). The freestream Mach number for
this case was set to 0.1, and the Reynolds number based on chord was taken to be
2.88 × 105. The body force for simulating the plasma actuator was obtained with
Model 1. Both steady actuation and unsteady actuation were considered to control
the turbulent separation. With a single steady actuator the size of the separation
bubble was reduced significantly, producing nearly a 20% reduction in the stream-
wise length of the separation region. The reduction in the separation region was
slightly less with unsteady actuation. As pointed out by He, this lower effective-
ness of the unsteady actuation is not consistent with what has been observed for
laminar separation control.
Gaitonde et al. (2005) investigated the effect of radio-frequency asymmetric
plasma actuators on the control of low-speed flow (M = 0.1) past a stalled NACA
0015 airfoil. Since the airfoil was stalled, the flow on the upper airfoil surface was
highly separated and unsteady. Such a flow represents a formidable challenge for
a flow control technique. The airfoil was at 15 deg angle of attack, and the flow
Reynolds number was 4.5 × 104. Simulations were performed for a wing with a
spanwise extent of 0.2 chords. The outer boundary of the domain was located at
30 chords from the wing. At the wing surface, no-slip and isothermal boundary
conditions were applied, and in the far field free-stream conditions were imposed.
Periodicity conditions were applied in the spanwise direction. For all simulations
a mesh with an O-H topology consisting of 308 × 75 × 145 points in the stream-
wise, spanwise, and body normal directions, respectively, was used. Spatial deriva-
tives of the RANS equations were approximated with a sixth-order compact
differencing method (Gaitonde et al., 1999; Visbal and Gaitonde, 2001). Spurious
frequencies in the solution were removed by a 10th order low-pass filter. To rep-
resent the effect of a plasma actuator on the flow, Gaitonde et al. computed a body
force. They determined the force with Model 2, and the parameters qf , Dt, a, rc,
and db in Eq. (83) were obtained from Shyy et al. (2002). The frequency of the
applied ac voltage was 3 kHz and the amplitude, which determines the electric
field strength and thus the body force magnitude, was varied. With Model 2 the
magnitude of E varies linearly, diminishing from a peak value near the exposed
electrode of the actuator until its breakdown value is reached. Four different ori-
entations of the body force were considered. When the body force vector was
primarily aligned with a surface tangential vector, a wall jet was created that
essentially eliminated the extensive flow separation on the airfoil upper surface.
The flow control based on plasma actuators produced nearly steady flow, with
only some low-frequency oscillations in the wake of the wing.

B. Polymers for Drag Reduction


In recent years many experiments and numerical simulations have been
conducted to try to understand the mechanisms of drag reduction by a polymer
MODELING AND SIMULATION 223

solution injected into a turbulent boundary layer. The high degree of effectiveness
that polymer solutions have in reducing drag has encouraged strong interest in
understanding the physics of the interaction of the polymer with the turbulence of
the fluid. Such understanding could possibly lead to the development of practical
active flow control devices to significantly reduce the drag of flight vehicles. In
the first part of this section we attempt to identify what is currently understood
regarding drag reduction due to polymer addition in small amounts. Then, after
presenting the basic equations for incompressible viscoelastic flows, we briefly
discuss some of the recent simulations for these flows.
Although much progress has been made in extending the theoretical foundation
of polymeric effects on drag (see, for example, Sreenivasan and White, 2000),
there is still an incomplete theory of the turbulent drag reduction process (i.e.,
reduction of the turbulence energy losses) when small amounts of certain poly-
mers are added to a fluid. The two primary theories to explain the drag reduction
came from Lumley (1969) and Tabor and de Gennes (1986). Lumley relies upon
the time criterion in his theory. The theory states that a polymer only has an effect
on the flow when the characteristic relaxation time of the polymer solution is
longer than a relevant turbulent time scale of the flow. He concludes that the drag
reduction is caused by substantial polymer stretching. The theory proposed by
Tabor and de Gennes is called the elastic theory, and it states that the stored elastic
energy of the polymers is responsible for drag reduction. Their paper also indi-
cates that polymers create an intervention in the turbulence energy cascade process
at the level of small scales. These two theories have provided focal points from
which many of the advancements in the understanding of the drag reduction pro-
cess due to polymers have emerged.
In a recent paper by Bonn et al. (2005) two principal questions are posed regard-
ing polymeric drag reduction. These questions concern the following: 1) where the
polymers intervene in the turbulent system (i.e., energy cascade process), 2) how
the polymers modify the flowfield. Regarding the first issue, it is generally accepted
that the polymers enter the turbulent flow system in the near wall region of the
boundary layer. Moreover, a recently performed experiment by Bonn et al. sup-
ports the conclusion that drag reduction is a boundary-layer effect. As to the modi-
fication of the flowfield by the polymeric addition, the evidence suggests that drag
reduction is due to the increase of the elongational viscosity, which is defined by
Bird et al. (1987) to be the resistance of a fluid to an elongational flow. If a small
amount of polymers is added to the fluid, the shear viscosity is not significantly
affected. In the paper by Wagner et al. (2003) experimental proof is given of the
connection between the elongational viscosity of a polymer solution and drag
reduction. In addition, Wagner and colleagues have demonstrated experimentally
that drag reduction increases with increasing polymer chain flexibility.
To investigate the polymer–turbulence interactions, higher-order (e.g., compact
difference, spectral) methods have generally been applied in direct numerical
simulations. Some simulations have also been performed with second-order
central difference schemes (e.g., Dubief and Lele, 2001). Due to the presence of
the polymers, the Newtonian stress tensor in the Navier–Stokes equations is aug-
mented by a non-Newtonian contribution. In general, a polymer-induced body
force also appears in the equations. The effects of the polymers in the fluid are
modeled. Both constitutive and kinetic polymer models, which are briefly
224 C. L. RUMSEY AND R. C. SWANSON

discussed by Joslin (2002), have been considered in the literature. Primarily, chan-
nel flows have been considered for investigating the physical aspects of viscoelas-
tic flows with polymers and numerical algorithms for solving such flows. However,
other types of flows such as isotropic turbulent flow and shear-driven turbulent
flow have also been used.
By simplifying Eq. (23) for incompressible flows and appending an additional
term due to the divergence of the polymeric stress tensor, we can write the nondi-
mensional momentum equations for a viscoelastic fluid as

∂ui ∂u ∂P b ∂ 2 ui 1 - b ∂t ikp
+ uk i = - + + (84)
∂t ∂x k ∂ xi Re ∂xk ∂xk Re ∂xk

where b is the ratio of the solvent viscosity ms to the mixture viscosity m, and
t ikp is the polymeric stress tensor. Since typical length scales associated with the
polymer molecules are much smaller than those for the turbulence, the effect of
the polymers on the flowfield must be modeled when applying standard numerical
techniques. One way to model the evolution of polymers is to use a bead-spring
model (often referred to as dumbbell model). With this type of model, multiple
massless beads are connected by springs to represent polymer chains, and each
pair of beads connected by a spring is called a dumbbell. The forces exerted on
each dumbbell include a hydrodynamic force due to the flow, a Brownian motion
force, and a spring force. A simple bead-spring model that is frequently used is the
Finitely Extensible Nonlinear Elastic (FENE) model. This model can be extended
to the continuum level by introducing the Peterlin approximation (Peterlin, 1961).
It requires the solution for an orientation (conformation) tensor

Cij = ·rirjÒ (85)


which represents the average orientation of the polymer chains at each point in the
fluid, with r being the separation vector between two beads. The transport equa-
tion for the FENE-P model that governs the conformation tensor can be written as

∂Cij ∂Cij ∂ui ∂u j


+ uk = Ckj + Cik - t ijp (86)
∂t ∂x k ∂x k ∂x k

The polymeric stress tensor is related nonlinearly to the conformation tensor and
is defined by
1 ( fC - d )
t pij = ___ (87)
ij ij
We
where We is the Weissenberg number, the ratio of the polymer time scale to the
flow time scales, and f denotes the Peterlin function, which is given by
L 2
f = _______ (88)
L2 - Ckk
The parameter L denotes the maximum polymer extension. Note that the trace of
the conformation tensor is the square of the bead separation distance, which must
always be less than the square of L.
MODELING AND SIMULATION 225

The FENE-P model is frequently used because it reproduces the effects of the
polymers on the wall turbulence characteristic length scales. It should be men-
tioned that numerical difficulties were experienced in many of the initial attempts
to use the FENE-P model with DNS for turbulent flow. Then, Sureshkumar and
Beris (1995) determined that the problem was caused by the growth of Hadamard
instabilities when solving the transport equation for the conformation tensor. To
circumvent this difficulty they introduced an artificial stress diffusivity to damp
the instabilities. The numerical difficulties encountered when solving the equa-
tions of motion for a viscoelastic fluid have also been considered by Vaithianathan
and Collins (2003). They proposed two numerical algorithms to remove some of
these problems. First, they modified the conformation tensor (Cij) to preclude
the polymer extension exceeding the finite extensible length. Second, they intro-
duced two possible matrix decompositions that allow reconstruction of the
conformation tensor so as to ensure positive definiteness, which must be main-
tained at all times for the calculation to remain stable. One important numerical
issue not addressed by their algorithms is the convection of the conformation
tensor. Since the standard FENE-P model does not contain a diffusive term,
there is no physical mechanism for preventing sharp fronts that can occur in the
conformation tensor due to turbulent convection. Based on their investigations,
Vaithianathan and Collins concluded that this appears to be the limitation of the
FENE-P model. In effect, this is precisely the limitation addressed by
Sureshkumar and Beris when they added an artificial diffusive term to the trans-
port equation for Cij.
With the stabilization of the numerical method for solving the equation govern-
ing Cij, Beris and his coworkers produced a series of papers (Beris and
Dimitropoulos, 1999; Dimitropoulos et al., 1998; Sureshkumar et al., 1997) con-
cerning drag reduction in a turbulent channel flow due to the addition of dilute
polymer solutions. The important findings from these efforts are summarized and
discussed by Beris et al. (2000). For example, with DNS they demonstrated that
the onset of drag reduction occurs once the relaxation time of the polymer becomes
comparable to a characteristic time scale of the turbulence (i.e., large eddy turn-
over time). Their results also provided evidence of a decreased eddy activity due
to the increase in elongational (extensional) viscosity. Some additional examples
of simulations with the FENE-P model are given in the papers by Brasseur et al.
(2005), De Angelis et al. (2002), and Dubief and Lele (2001).

C. Parameter Optimization
Catalano et al. (2002) optimized the performance parameters of one and two
synthetic jets with the objective of minimizing the drag coefficient of a circular
cylinder with flow control. They used a response-surface optimization method,
which is a direct technique since it requires no gradient information to minimize
the objective function. They combined the optimization method with a DNS/
LES approach to determine flow solutions with the synthetic jet actuators. An
energy conserving Navier–Stokes solver (Choi, 1993) with a hybrid finite dif-
ference and spectral scheme (Mittal and Moin, 1997) was used for numerical
flow simulations. To advance a solution in time, Catalano and colleagues applied
a fractional step method along with the Crank–Nicolson scheme for the viscous
226 C. L. RUMSEY AND R. C. SWANSON

terms and a third-order Runge–Kutta algorithm for the convective terms. In their
computations synthetic jets were modeled by imposing the velocity normal to
the surface as

Ê V tˆ
V j = g(q j )(V j ) A sin Á 2p f j • ˜ (89)
Ë D¯

where g(qj) is a top-hat function (Rizzetta et al., 1999), V• is the free-stream


velocity, qj is the jet location, D is the cylinder diameter, and the frequency fj = kj fs
with fs being the natural shedding frequency of the flow. The parameters (Vj)A, kj,
and qj were determined with the optimization procedure and 2-D simulations on a
C-type mesh with 201 × 60 points. With the optimized parameters, Catalano et al.
(2002) obtained LES solutions for flow over the cylinder with spanwise extent of
four diameters. They used the dynamic subgrid scale model of Germano et al.
(1991) in the computations, and the grid consisted of 401 × 120 × 49 points. At a
Reynolds number of 500 they obtained a modest reduction of drag coefficient of
8–13%, when they applied one and two synthetic jets for flow control. When the
Re was increased to 3900 they observed no drag reduction. These results are not
totally surprising since the control parameters were determined with 2-D simula-
tions, and the dynamics of the 3-D computations can be quite different. In addi-
tion, the 2-D simulations were made with the subgrid scale model switched off,
which could affect the synthetic jet parameters.

D. Neural Network
One simplified approach for computing the effects of a synthetic jet on a flow-
field involves modeling the jet using a neural network trained by lumped deter-
ministic source terms. Filz et al. (2003) considered this approach for modeling
directed synthetic jets. In order to provide the necessary data for training the
neural network, they computed several solutions to the unsteady 2-D RANS equa-
tions for a synthetic jet immersed in a low-speed flow. They used a solver with
second-order accuracy in space, fourth-order accuracy in time, and low-speed
preconditioning. To represent the effects of turbulence they applied the wall-
distance-free k-e turbulence model of Goldberg et al. (1998). The required source
terms were computed by forming the time-averaged dependent variables from the
unsteady solutions, and then forming residuals of the flow equations. Thus, these
source terms included the unsteady effects and provided the necessary data for
training the neural network. Filz et al. (2003) based the neural network on the fast
and accurate Levenberg–Marquardt back propagation algorithm (Levenberg,
1944; Marquardt, 1963). They stored the five source terms for each grid cell and
trained the neural network to determine a new set of source terms as a function of
location and jet orifice orientation angle for a fixed external flow Mach number.
To verify the source terms of a flow not used in training the neural network, Filz
et al. (2003) performed steady-state computations with the source terms generated
by the neural network and compared the results with the time-averaged solution
for that flow. They demonstrated that the neural network source terms allowed
accurate reproduction of the time-averaged physics.
MODELING AND SIMULATION 227

E. Control Theory
There are many possible objectives in flow control, such as drag reduction,
separation control, and enhanced mixing, that can be achieved through the use of
control theory. Many of these objectives fall under the general category of turbu-
lence control. Various methods utilizing control theory have been proposed for
flow control, and several review articles (i.e., Bewley, 2001; Lumley and Blossey,
1998; Moin and Bewley, 1994) discuss these methods. There are several possible
control strategies, including open-loop, closed-loop (i.e., feedback control), and
optimal control. In open-loop control the actuator parameters are established in
the design stage, and they remain fixed even if there are changes in the flow state.
A closed-loop control uses data from the flow state along with a model of the state
to devise changes in the actuator parameters to achieve the control objective. With
this type of control, sensors are used to make state measurements. Optimal control
methodologies can be applied to both open-loop and closed-loop control. It can be
considered a general framework for flow control of nonlinear systems. In the
simplest sense optimal control is a minimization process based on a family of
desired controls and an objective functional. Moreover, a set of differential equa-
tions and their adjoints are considered. Then optimization is performed on either
the continuous problem (followed by discretization) or on the discrete problem. A
general discussion of these different types of control is given in the paper by
Collis et al. (2004). In this section we discuss some applications of these control
methods. Because examples of open-loop control have been presented in other
sections, the focus here is on closed-loop and optimal control. An extension of
optimal control is also briefly described.
In a closed-loop control some measured quantity is converted to a control
input by a controller. Frequently the controller is designed to have a physics-
based strategy for converting measured input into flow control input for an actua-
tor. One physics-based control strategy is called opposition control, a type of
feedback control that has been applied to near-wall turbulence control. Opposition
control involves a combination of surface suction and blowing to oppose the
near-wall velocity of streamwise turbulence structures. Choi et al. (1994) and
Hammond et al. (1998) used opposition control in conjunction with DNS to
demonstrate for a turbulence Reynolds number Ret = 180 and sensing planes
located at y+ = 10 and y+ = 15, respectively, drag reduction of 20 and 25%. They
also showed that drag increases occur if the sensing plane is not placed suffi-
ciently close to the wall.
We now consider two examples of active flow control in which optimal control
is applied. First, Joslin et al. (1995a, 1997) developed a general optimal control
method that can be used for a variety of flow control problems, including separa-
tion control and transition suppression. In their method they coupled the time-
dependent incompressible Navier–Stokes equations with the adjoint Navier–Stokes
equations and optimality conditions in order to determine the optimal states (i.e.,
unsteady flowfields and controls). They used DNS to solve the flow equations. The
flow equations were integrated forward in time and the adjoint equations backward
in time. They applied their method to the problem of transition suppression, and
demonstrated that instabilities can be suppressed without a priori knowledge of the
disturbance. This represented a departure from other control methods.
228 C. L. RUMSEY AND R. C. SWANSON

Two especially important issues for these types of optimal control applications
are boundary conditions and computational effort. First, the boundary conditions
require careful consideration, especially at the outflow and inflow boundaries,
where wave reflections can easily occur with inappropriate treatment. Joslin et al.
avoided this problem by selecting a short enough time interval so that the waves
did not reach the boundary. An alternative approach is to implement appropriate
buffer conditions (see Dobrinsky and Collis, 2000). Second, the storage require-
ments and computational expense can easily become prohibitive for 3-D problems.
The computational effort can be reduced by limiting the optimization time interval
and/or using reduced-order modeling (see Collis et al., 2004).
The second example of using optimal control involves feedback control of flow
separation. Choi et al. (1999) considered a feedback control method and two
boundary control strategies for the time-dependent incompressible flow over a
backward-facing step. Their primary objective was to develop a robust control
technique to reduce the size of the separation region behind the step. On the sur-
face behind the step they placed a flow sensor, and at the start of the step they
applied a control function representing blowing and suction. Choi et al. defined a
two-part cost functional, one part penalizing a negative wall velocity gradient in
the sensor area containing the flow reattachment point, and the other part weight-
ing the cost for the control. They solved the RANS equations and the correspond-
ing adjoint equations with a semi-implicit scheme. A gradient algorithm was used
to reduce the cost functional. With this approach they demonstrated significant
reduction in the separation region. Furthermore, they developed a general theoreti-
cal framework to obtain feedback control laws for the time-dependent, incom-
pressible Navier–Stokes equations using Lagrangian techniques.
Optimal control theory can be generalized to what is called robust control theory.
Bewley et al. (2000) proposed a general framework for robust control in fluid
mechanics. Their theory has been developed for problems in which the Navier–
Stokes equations are being solved. To solve the robust control problem, Bewley
et al. proposed an algorithm using repeated computations of an adjoint field. Since
the computation of the adjoint field is only as difficult as the computation of the
flow solution, the numerical adjoint problem is tractable whenever the numerical
flow problem is tractable. One advantage of their algorithm is that it involves
computations O(N) (N is the number of grid points used to solve the flow prob-
lem), whereas control methods based on Riccati equations or Hamilton–Jacobi–
Bellman formulations require O(N2) computations. They have used their algorithm
for turbulence control problems.

V. Summary and Conclusions


Various CFD methodologies typically used for solving active flow control
applications have been presented. Most applications assume the working fluid is a
single-component perfect gas, and the equations have been shown here from that
perspective. Through searching the flow control literature, we have attempted to
glean a representative sample of numerical considerations and issues encountered.
The methods of DNS, LES, and RANS have been described in detail, and other
methods such as blended RANS-LES and techniques such as immersed boundary
methods have been summarized.
MODELING AND SIMULATION 229

True DNS has been shown to be a very powerful but impractical tool for high
Reynolds number applications using today’s computers. However, under-resolved
DNS is sometimes used, and may yield valuable predictive information for certain
aspects of many flow control problems. Although still not routinely used, LES has
been significantly enabled recently because of continued increases in computer
capabilities. It appears to hold great promise for computing these flows. Finally,
RANS methods have served as the backbone for flow control computations, used
routinely for a wide range of applications. Naturally, the fundamental tenet of
RANS—that the actions of all random turbulence fluctuations can be described in
a mean sense using models for unknown tensor correlations—has been questioned
for its applicability to many flow situations. Turbulence models are routinely
blamed for poor agreement with experiment, but we have shown that other model-
ing uncertainties (e.g., inconsistent boundary conditions, running 2-D for an
inherently 3-D problem, etc.) can also often share much of the blame.
Two recent CFD validation workshops were also summarized. These exercises
have been instrumental in pointing out strengths and weaknesses in various meth-
odologies for solving flow control problems. In particular, RANS models have
been shown to give mixed results, sometimes agreeing well with experiment and
sometimes not, depending on the case. Coanda wall jet flows have been shown to
be very sensitive not only to turbulence models but also to numerical parameters,
and therefore are difficult to compute reliably. In regions of significant separation,
the turbulence levels from RANS models tends to be underpredicted, giving reat-
tachment locations too far downstream. It appears that allowing the larger eddies
in the separated shear layer to develop and interact on their own through LES,
blended RANS-LES, or DNS type methods can significantly improve this defi-
ciency. These workshops have also demonstrated some of the difficulties associ-
ated with taking and comparing against flow-control related experimental data,
and have pointed to the need for continued high-quality experimentation in this
area. In particular, the need for extremely well documented boundary condition
information near jet and suction plenum exits has been described.
We have also described computational methods that are being used to explore
what we perceive as exciting advanced control strategies. Many of these strategies
are still under active development, and have not seen wide practical applications
to date. As such, the computational methods currently used to model them are also
in flux. No doubt the next decade will show rapid advances in modeling and simu-
lation, and new techniques will emerge that will help drive this dynamic discipline
forward.

Acknowledgments
The authors would like to thank Donghyun You of the Center for Turbulence
Research and Veer Vatsa of NASA Langley Research Center for providing results
from their computations.
Chapter 8

Fixed Wing Airfoil Applications

Avraham Seifert*
Tel-Aviv University, Tel-Aviv, Israel
and
Carl P. Tilmann†
Air Force Research Laboratory, Wright–Patterson Air Force Base
Dayton, Ohio

I. Overview
The purpose of this chapter is to provide a summary of past, present and future
airfoil boundary-layer separation control studies as an example of active flow control
application. The practical importance of boundary-layer separation control has been
identified already by Prandtl (1904), when he introduced the boundary layer concept
and characterized its failure, i.e., separation. Avoiding boundary-layer separation
would bring us closer to the ideal flow conditions with enhanced system perfor-
mance at lower energetic cost. Boundary-layer separation control heavily relies on
the performance of flow control actuators and sensors, reviewed in Chapter 6.
Preliminary pioneering separation control efforts will be reviewed first. These
will include shear-layer and airfoil transition and separation control studies. The
state of the art in airfoil separation control will then be critically reviewed. This
will lead to recommendations of required progress that has to be made in order to
explore the great potential of active separation control as an enabling technology
for future aeronautical, transportation, and many other fluid related systems.
The first system study to consider the application of unsteady active flow control
(AFC) for transport aircraft was conducted in the late 1990s by McLean et al.
(1999). They considered multiple applications of AFC and concluded that simpli-
fying the high-lift system while maintaining Cl,max, is the most promising applica-
tion. It could lead to lower part count, lighter structure, cost reduction and most

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Associate Professor, Mechanical Engineering, School of Mechanical Engineering. Associate

Fellow AIAA.
†Senior Technical Advisor, Aerodynamic Configuration Branch. Associate Fellow AIAA.

231
232 A. SEIFERT AND C. P. TILMANN

importantly (current authors’ interpretation) to 1–2% cruise drag reduction due to


the elimination of all external flap positioning actuators. These benefits could
result from effective and robust boundary-layer separation control. However, a
decade later, its conclusions are far from being realized. Why? Two main reasons
can be identified. First, the investigators linearly extrapolated from existing low
Reynolds number results both to high Re and to higher flap deflections with
linearly increasing lift increment, with respect to what is known from uncontrolled
deflected flap data. Second, at the time, published work did not include a compu-
tational fluid dynamics (CFD) design tool as an enabling technology, so ad-hoc
assumptions had to be made. Until a computational design tool for unsteady
boundary-layer separation control becomes available, flow control would largely
remain an art. Hopefully, once such a tool was available the artistic magic would
not disappear, since this complex field of study requires immense innovation.

II. History and Background


Whereas Chapters 1 and 2 have thoroughly described the physics and provided
a comprehensive introduction to AFC, here, we briefly highlight key aspects of
history related to fixed wing separation control.
Boundary-layer control (BLC) research dates back to the turn of the twentieth
century, when Prandtl (1904) introduced the concept of the boundary layer, its
failure (i.e., separation), and a possible remedy (e.g., removal of the near wall
fluid by slot suction). Half a century passed before Schubauer and Klebanoff
(1956) conducted the first AFC experiment when they artificially triggered
Tollmien–Schubauer waves in a laminar boundary layer. An additional 20 years
passed before—in the mid 1960s—Collins and Zelenevitz (1975) conducted the
first separation control experiment, using boundary-layer transition promotion by
sound emanating from the tunnel walls that eventually led to separation control.
In the abovementioned and other subsequent studies, transition promotion and
separation control were mixed. Moreover, a limited range of Reynolds numbers
was investigated. The airfoil experiments of Seifert et al. (1996) and Seifert and
Pack (1999) resolved the frequency scaling issue and demonstrated active
boundary-layer separation control technology at chord Reynolds numbers ranging
from 100,000 to 30,000,000. The wall-mounted hump experiments by Seifert and
Pack (2002) demonstrated the validity of previous findings related to frequency
scaling in a fully turbulent environment and provided benchmark data for CFD
validation. The above fundamental experiments paved the way to many studies
aimed at applying this knowledge to specific geometries, on the way to making
this technology useful for real-world applications.

III. Examples from the Present State of the Art


The purpose of this section is not to provide an unbiased review of the state of
the art of active boundary-layer separation control, but to illustrate several appli-
cations of separation control using examples in order to enrich the knowledge of
the readership and be specific in describing the principles.
The chosen examples of active separation control (ASC) are related to rather
short, but extremely important segments of the flight—take-off and landing.
FIXED WING AIRFOIL APPLICATIONS 233

High-lift is required for both of these segments of the flight due to the low aircraft
speed. While energy consumption during these segments is usually not of great
importance for the overall mission energy considerations, it would be disadvanta-
geous if the ASC system were seen to add weight, raise compatibility or reliability
issues, or have adverse environmental implications.

A. High-Lift—NASA-TAU Simplified High-Lift System


The purpose of this series of simplified high-lift system experiments (Pack-
Melton et al., 2004, 2005, 2006, 2007) was to explore the prospects of eliminating
all but simply hinged leading and trailing edge flaps, while controlling boundary-
layer separation on a supercritical airfoil using multiple periodic excitation slots.
Excitation was provided by three, independently controlled, self-contained piezo-
electric actuators. Low-frequency excitation was generated through amplitude
modulation (AM) of the high-frequency carrier wave, the actuators’ resonant
frequencies. The low-frequency excitation was used to operate at the optimal
Strouhal numbers (Seifert et al., 1996) of order unity, rather at the actuators’ reso-
nant frequency (order of 1–2 kHz) at Strouhal numbers greater than 10. This high
natural frequency of the actuators was found to be inefficient for the present
configuration and flow conditions, so AM was used. The primary challenge of
these studies was to effectively combine several excitation sources. The use of
AM to generate low-frequency excitation from the O(1 kHz) frequencies of the
piezo ZMF actuators was first suggested by Glezer and co-workers (Amitay et al.,
2001; Wiltse and Glezer, 1993). It was later demonstrated by Yehoshua and Seifert
(2006a) to be related to the nonlinearity of the cross-flow response to the excita-
tion (following Ingard and Ising, 1967).
Flow control research using steady momentum injection into a high-lift system
separating boundary layer dates back to the 1930s (Lachmann, 1961, p. 26).
Additional interest was spurred in the 1950s by the use of the gas turbine engine
which could provide a source of compressed air. Poisson–Quinton and Lapage
(1961) showed that separation could be effectively controlled using wall-tangential
steady momentum transfer (blowing) but that the momentum requirement was not
insignificant (e.g., between 3–5% Cm ). The use of periodic excitation for separa-
tion control on the simply hinged high-lift system should significantly reduce the
momentum and power requirements compared to that of steady mass transfer
(Seifert et al., 1996), as shown in Fig. 26. In addition, research using pulsed exci-
tation has shown that the momentum requirements and associated power expendi-
ture can be reduced further by varying the duty cycle of the excitation significantly
below 50% (Margalit et al., 2002).
The maximum lift of the airfoil was enhanced by a two-stage mechanism. The
leading edge (LE) actuator (marked 1 in Fig. 1) maintained attached flow from the
actuator to the trailing-edge (TE) flap-shoulder region (marked 2 and 3 in Fig. 1).
Activating the TE flap-shoulder actuator in a manner that properly interacts with
the vortex shedding process (therefore requiring low-frequency excitation) provided
an additional lift increment. This interaction enhanced the suction level upstream
of the flap region actuators, while significantly enhancing vortex-shedding magni-
tude downstream, but at a frequency roughly 50% higher than the natural shedding
frequency of the separated flap.
234 A. SEIFERT AND C. P. TILMANN

Fig. 1 The simplified high-lift system modified EET airfoil used by Pack et al. (2004,
2005, 2006, 2007).

Experiments conducted at both low (Seifert et al., 1996) and high (Seifert and
Pack, 1999, 2002) Reynolds numbers have shown that periodic excitation is effec-
tive at controlling separation as well as efficient in terms of the required control
authority. This information, combined with that of a system study by McLean
et al. (1999), indicating the possibility of significant payoffs such as net airplane
cost, weight, and cruise drag reductions, has motivated research paving the way to
the application of ASC to a simplified high-lift system. The purpose of these stud-
ies was to explore ways to simplify current multi-element high-lift systems and
airfoils that use slots (Lin et al., 1994) and the Fowler effect (i.e., increasing lift
by not only deflecting the flap for larger lift coefficient, but also increasing the
surface by moving the flap downstream) to generate high-lift and vortex genera-
tors to further delay and manipulate separation. The chosen design completely
eliminates hinges and positioning actuators which need to be external to the airfoil
contour, as well as passive slots for energizing the boundary layers. These slots,
especially the associated additional trailing edges, are known to be a significant
noise source. All hinges and positioning actuators in the simplified high-lift con-
figuration are internal, and thus eliminate the associated parasite drag at cruise.
The LE flap is used to increase CL,max due to increased circulation and prevention
of laminar LE separation at high incidence. Zero-mass-flux (ZMF) periodic exci-
tation, directed downstream at a shallow angle to the local surface, is applied at
locations that are prone to separation, i.e., the LE and TE flap shoulders.
The results obtained when applying periodic excitation at the LE flap shoulder
of this airfoil were reported in a series of publications (Pack-Melton et al., 2005).
High-frequency, f, periodic excitation, typical of the piezoelectric actuators, was
applied at the LE flap shoulder, delayed stall, and increased the maximum lift
coefficient (CL,max) by 10–15%, at low TE flap deflections (Fig. 2). It was shown
that low-frequency AM could be used to achieve similar benefits in aerodynamic
performance and required significantly less power (Fig. 3). The excitation momen-
tum coefficient is not necessarily a suitable scaling parameter for the amplitude
effect and this issue is still open, as will be discussed later.
The effect of applying excitation at various locations along the TE flap was also
examined (Pack-Melton et al., 2006; Fig. 4). Upto then, no study had been con-
ducted to examine so thoroughly the sensitivity to the slot location on the flap.
More importantly, no effort had been made to combine several excitation sources
on the flap, with the aim of utilizing constructive interference for enhanced control
authority. It was found that the optimal flap setting for a given excitation slot
location was very sensitive and changed significantly (Seifert and Pack, 2003a;
Fig. 5). This was attributed to the small surface radius of curvature at the forward
part of the TE flap. Low-frequency AM of the high-frequency excitation reduced
the momentum requirements for a given lift increment by a factor of two to three
FIXED WING AIRFOIL APPLICATIONS 235

LIVE GRAPH
Click here to view

Fig. 2 The effect of LE region excitation on the simplified high-lift system, using low
TE flap deflection, as a function of the LE flap deflection.

LIVE GRAPH
Click here to view

Fig. 3 The effect of pure sine and AM excitation on lift vs excitation magnitude
(Rec = 0.24 × 106, a = 0 deg, ds = -25 deg).

Fig. 4 Excitation locations on the flap, corresponding to the region marked 3 in


Fig. 1 (EET airfoil above).
236 A. SEIFERT AND C. P. TILMANN

LIVE GRAPH
Click here to view

Fig. 5 The effectiveness of the different slots on the flap of the simplified high-lift
system, Re = 0.4 × 106.

and also reduced the sensitivity of the lift increment due to AFC to flap setting and
slot location. It was hypothesized that the longer wavelength is less sensitive than
a shorter wavelength to local surface curvature. The momentum requirements for
controlling separation on the TE flap were found to be significantly higher than
those required for control of separation near the LE flap shoulder and also as
compared to high Reynolds number experiments performed on a NACA 0015
airfoil (Seifert and Pack, 1999). This aspect certainly requires further study.
To increase the momentum and vorticity flux available for controlling separa-
tion on the TE flap, the effect of combining the excitation from an actuator just
upstream of the TE flap (i.e., on the most aft region of the main element) with the
excitation ejected from an actuator just downstream of the TE flap shoulder (i.e.,
on the forward part of the TE flap) were examined. The effects of duty cycle and
phase angle (Fig. 6) between the two actuators were studied as well, with the aim
of maximizing control authority while minimizing energy expenditure. Finally
the effect of combining the TE flap excitation with the excitation emanating from
the LE flap actuator was performed.
It was demonstrated that pulsed-modulated signals from two neighboring slots
can favorably interact to increase lift. Phase sensitivity of the modulation fre-
quency was measured, even though the excitation was synthesized from the high-
frequency carrier wave (Fig. 6). It was found that the additional activation of an
LE flap-shoulder actuator slightly increased the most effective excitation fre-
quency at the TE flap shoulder. It was also found that the phase sensitivity, which
enables enhanced performance due to favorable interaction between two TE flap-
shoulder actuators, disappeared at incidences corresponding to maximum lift.
FIXED WING AIRFOIL APPLICATIONS 237

LIVE GRAPH
Click here to view

Fig. 6 Sensitivity to excitation phase lag between the TE actuator (in region 2 in
Fig. 1) and the flap slot #3 actuator (see Fig. 4) (FM = 40 Hz, F +M,TE = 0.52, F +M,#3 = 0.47,
Rec = 0.24 × 106, ds = -25 deg, d f = 20 deg).

Finally, ZMF periodic excitation was introduced from three locations on a


simplified high-lift system (Pack-Melton et al., 2004). Pulsed modulation was
again used to generate low-frequency inputs and study relative phase effect when
combining neighboring actuators, each resonating at a slightly different frequency
but all at around 1 kHz. Excitation was introduced from an actuator placed just
upstream of the TE flap shoulder to increase the momentum available for control-
ling flow separation over the TE flap shoulder at large TE flap deflections.
The TE actuator (marked 2 in the airfoil Fig. 1 above) was more efficient for lift
increment when the high-frequency excitation produced by the actuator at its reso-
nance frequency was modulated at low frequency. The momentum required for
similar lift increment could be reduced by as much as 50% when AM was used.
By using burst modulation (BM) and varying the duty cycle, the efficiency of the
TE actuator could be further improved over that of AM excitation.
Combining the TE and flap excitations increased the maximum lift over that
produced by either actuator operating separately. Using pulsed modulation (indi-
cated BM on the figures) it was found that the phase angle of the modulating
waveform between the input signals had a significant but complex effect on both
lift and form-drag. Figure 6 above shows the lift dependence on the relative phase
between the actuators. The maximum increment in lift occurred for a phase angle
in the range of ±30 deg. An additional maximum lift increment of about 15% was
achieved when combining the LE and flap actuators (see Fig. 7). Further research
is required to interpret some of the complex behavior observed. However, detailed
particle image velocimetry data revealed that the upstream effect of the optimal
combination of flap shoulder region dual-interacting excitations results from the
generation, interaction, and shedding of “locked” vortices (see Fig. 8).
238 A. SEIFERT AND C. P. TILMANN

LIVE GRAPH
Click here to view

Fig. 7 Improvement of the lifting characteristics of the EET simplified high-lift


system due to multiple slots excitation (Rec = 0.24 × 106, ds = -25 deg, dF = 20 deg).

B. High-Lift—ADVINT for the Boeing Tilt-Wing


The Adaptive Flow Control Vehicle Integrated Technologies (ADVINT)
program (Smith et al., 2006) explored the use of AFC to enable simplified TE
high-lift systems for Super Short Take-Off and Landing (SSTOL) aircraft. This
program focused on the maturation of an integrated high-lift system incorporating
electromagnetic ZMF jet actuators for breakthrough aerodynamic performance.
The target platform for this technology was the Boeing tilt-wing (BTW) SSTOL
transport concept (Manley and von Klein, 2002), but the performance goals for
the simplified flap system were derived from the LTV XC-142 tilt-wing technol-
ogy demonstrator (Goodson, 1966; Dausman, 1990), which had a much more
complex three-element flap system. The BTW concept uses flow control to achieve
equivalent levels of lift by controlling separation over a simple flap—enabling a
reduction in the thrust required for takeoff, and a reduction in the wing tilt angle
required for landing.
A comprehensive series of 2-D and 3-D experiments were performed to opti-
mize the AFC system for this application, and the final demonstration was on an
FIXED WING AIRFOIL APPLICATIONS 239

LIVE GRAPH
Click here to view

Fig. 8 The optimal combination of two flap actuators (shown as white lines).

11%-scale, powered semi-span model tested in the NASA Langley 14 × 22-ft


Subsonic Tunnel. In preparation for the final demonstration, a series of small-
scale experiments (5%) were conducted to aid in understanding the governing
AFC mechanisms, identifying the proper parameters to characterize and optimize
the performance of the AFC system, understanding the effects of wing sweep on
AFC effectiveness, and determining the scaling relationships needed to develop
AFC systems for large-scale test and full-scale applications. Other important
objectives of the small-scale experiments were to examine different airfoil con-
figurations to determine which geometries were most amenable to AFC, to com-
pare AFC performance in 2-D and 3-D configurations, to determine the optimal
AFC slot locations for different flap deflections, and to establish the expected
AFC amplitudes required to reach the takeoff and landing performance goals. In
addition, the electro-magnetic ZMF actuators to be used in the large-scale demon-
stration were greatly improved to deliver high jet velocities and increase reliability.
These investigations have been described in detail by Nagib et al. (2006, 2007a,
b) and Kiedaisch et al. (2006, 2007) and are summarized here.
Early experiments studied the effectiveness of AFC for a variety of LE and flap
configurations on a high-lift 2-D airfoil. The AFC method chosen for these initial
tests was ZMF oscillatory downstream forcing through a thin 2-D slot near the
flap shoulder. The apparatus which was used allowed a range of amplitudes, forc-
ing frequencies, and AFC slot locations to be explored, providing important guid-
ance for the design of the flow control system for the 11% model. The effects of
AFC introduced at the flap shoulder were manifested in the pressures across the
airfoil’s entire upper surface and on the LE slat (when there was one). With intense
enough forcing applied at the flap shoulder, AFC was effective at increasing the
240 A. SEIFERT AND C. P. TILMANN

circulation on the airfoil. The best AFC performance was observed at high peak
jet velocities, and at relatively low forcing frequencies, and the effect on lift was
proportional to the ratio of peak jet velocity to the free stream velocity (Kiedaisch
et al., 2006).
The 2-D airfoil model was then mounted on an adapter plate in a 3-D forward-
swept configuration to evaluate AFC performance in the presence of crossflow
and wing tip effects (Smith et al., 2006). As expected, the baseline 3-D section
lift coefficients were lower than for 2-D below the stall angle of attack. However,
the percentage change in section CL due to AFC was consistently greater for
the swept airfoil than it was for the 2-D case, and AFC effectiveness increased
near the tip of the wing. This is in agreement with earlier studies by Naveh et al.
(1998) and Seifert and Pack (2003b), where both offered scaling options between
the 2-D and 3-D flow conditions and demonstrated that modest sweep angles do
not alter the governing parameters and their optimal values with respect to the 2-D
knowledge.
Other forms of flow control including steady and pulsed blowing and suction
were also explored by Nagib et al. (2006). Results were obtained on the 2-D airfoil
configuration equipped with a simply hinged flap, and a LE slat (to prevent LE
separation at high flap loadings). These results were then compared to inviscid
airfoil computations (an idealized goal) and to the ZMF AFC results. At low
amplitudes, steady blowing and ZMF appeared to provide a similar performance
for equivalent jet velocity ratios. It was experimentally shown that steady blowing
with sufficient amplitude was capable of achieving—or even exceeding—inviscid
levels of performance. All the data indicate that while the applied AFC is capable
of controlling separation, improvements in the sectional lift coefficients do not
necessarily mean that the separation was eliminated—it may be that the circula-
tion is simply being enhanced. This implies that controlling circulation and sepa-
ration are intricately woven together, but both should be considered when
designing the AFC system.
The effectiveness of AFC while the flap experienced a dynamic deflection was
also investigated (Kiedaisch, 2007). This emulated the proposed concept of rap-
idly rotating a simple flap from a negative position (to maximize ground accelera-
tion) to full deflection thereby rapidly enhancing lift immediately before takeoff.
While these dynamic flap tests were limited in scope, they indicated that the lift
enhancement seen on a static simple flap is also achieved when the same control
is applied to a dynamically deflected simple flap—independent of flap deflection
rates of interest for this application.
In all the small-scale experiments described above, the actuation system was
external to the model, and the pressure fluctuations were routed through manifolds
into a plenum in the flap. A separate wall-mounted model, referred to as the
pseudo-flap, was used to develop and test the actuators to be used in the large-
scale testing, and to aid in the slot design. The pseudo-flap was designed to pro-
duce a pressure distribution similar to the 2-D flap’s upper surface, and was used
to explore various types of separation control, including steady, oscillatory, and
pulsed forms of suction and blowing. One principal difference between the flows
over the pseudo-flap and the airfoil is the ability to modify the global circulation
around the airfoil. This changes the flow near the control location, and increases
the pressure recovery that the AFC is required to manage. Like the airfoil model,
FIXED WING AIRFOIL APPLICATIONS 241

the location of separation on the pseudo-flap can move in response to the AFC
effects. This is different from most other “hump-like” wind tunnel models
(e.g., Greenblatt et al., 2005; Seifert and Pack, 2002), which were designed to
have a nearly fixed separation location to make the results more amenable for use
as an AFC validation case for CFD simulations. In agreement with prior research,
suction was found to be much more effective than blowing, in that it required
lower flow rates to be effective, and was less sensitive to slot location and orienta-
tion (Nagib et al., 2006).
Finally, the large-scale demonstration tests were conducted in the NASA
Langley Research Center’s 14 × 22-ft subsonic wind tunnel on a 10.75% scale
semi-span model of the BTW conceptual aircraft. Actuators were installed in all
flap sections across the span of the wing. Early testing focused on AFC system
integration and development, and on verifying predictions of the flow control
actuator parameters for maximum lift enhancement. The resulting force and
moment increments due to AFC were not significantly affected by variations in
actuator frequency in the neighborhood of 0.3 < F + < 1.5 (based on inboard flap
cord length and freestream velocity), but AFC effectiveness noticeably improved
as actuator amplitude was increased.
The objectives of the rest of the testing were to acquire vehicle-level performance
data, and to assess the impact of AFC on overall aircraft performance to include the
influences of propeller wash, wing tilt, and aircraft angle of attack. These were
the first AFC experiments to include unsteady propeller flow over a swept wing
section. For all but one run the actuators were operated at 90 Hz (F +  1–1.15)
with all actuators working in-phase. It is worth mentioning here that there is evi-
dence that operating segmented actuators out of phase can be beneficial for lift
increment, drag reduction and reduced power consumption (Timor et al., 2007).
The AFC performance improvement goal for the takeoff configuration was met
over almost the entire range of angles of attack evaluated, and the target for the
landing configuration was met over a substantial range of angles of attack (see
Fig. 9), producing significant lift improvements with virtually no changes to the
drag polars. The optimum actuator operating frequencies determined from the
2-D experiments were well suited to the 3-D configuration testing. Most signifi-
cantly, AFC was able to reduce the amount of thrust required for the take-off
maneuver and reduce the operational angle of attack by 2–5 deg at take-off and
landing, increasing the stall margin and reducing the required vehicle attitude or
wing tilt angle.

C. Drag Reduction of Streamlined and Bluff Bodies


In 2-D flow, drag results from skin friction and pressure or form-drag. The term
drag reduction via AFC to be (subsequently) used here, is generally related to the
reduction of form-drag. Reducing turbulent skin friction is a hard task, not
achieved to date via flow control. There is no known instability mechanism that
can make this process energy efficient.
Enhancing the skin friction of a marginally separated (e.g., with vanishing Cf),
boundary layer is an achievable task; however, its inherent sensitivity to any pos-
sible parameter variation requires a robust closed-loop AFC (CLAFC) system
with distributed arrays of sensors, actuators, controllers, and significant power
242 A. SEIFERT AND C. P. TILMANN

LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view

Fig. 9 Increments in CL due to ADVINT AFC: a) slats extended, take-off dF = 40 deg,


and b) landing dF = 50 deg. CT = thrust coefficient = thrust/qSREF (Smith et al.,
2006).

consumption. On the other hand, reducing the drag of massively separated bodies,
such as bluff bodies at relatively high-Re, is achievable via two mechanisms. The
first is by a delay of boundary-layer separation, enabling higher base pressure at
the lee-side of the body and therefore lower form-drag. The other mechanism is
wake-vortex-shedding manipulation. This mechanism requires CLAFC as well,
but it is simpler since the shedding is a globally unstable mechanism, highly
receptive to excitation and enables sensing the VSF, its amplitude and phase by
sensors located on the body itself (Stalnov et al., 2007).
FIXED WING AIRFOIL APPLICATIONS 243

D. Moment Generation—Low Incidence Control—Reattachment vs


Separating Attached Flows
The generation of control moments for aerodynamic steering of airplanes is
highly desirable from many practical points of view. For military vehicles, elimi-
nating the vertical control surfaces could lead to significant reduction of the radar
cross-section. The same task requires that other control surfaces should not be
used during critical segments of the mission. From a civilian point of view, reduc-
ing the size of the vertical stabilizer and rudder, dictated by engine-out perfor-
mance, could lead to significant cruise drag-reduction. The generation of
aerodynamic control moments using AFC is expected to be of significantly wider
bandwidth than deflecting conventional control surfaces, due to inherently higher
and prohibitive inertial loads created by the latter. Here also, however, marginally
separated baseline boundary layer makes this task achievable. With fully attached
boundary layer this task is possible only with O(1–10) control authority actuators.
While energy is not a consideration in this application due to the short periods
such a mechanism is designed to operate, available actuators are of either limited
control authority (e.g., piezofluidic of Yehoshua and Seifert, 2006b) or prohibi-
tively complex and potentially non-integrable, dangerous, expensive, and heavy,
such as spark-jet actuators (Cybyk et al., 2005) and combustion chamber actuators
(Cutler et al., 2005). Utilizing flow separation is also possible by mechanically
separating a boundary layer by a backwards facing step (DeSalvo and Glezer,
2006; Glezer, 2008) or by wall normal fluidic steady or ZMF jets, but at a momen-
tum ratio of at least O(1) (Chen et al., 2000).
A pioneering demonstration of roll control via AFC was provided by Seifert
et al. (1999). A half-scale Hunter UAV was instrumented with an oscillatory blow-
ing system that provided pulsed excitation over the flaps (see top view in Fig. 10).

Fig. 10 The use of an oscillatory blowing valve to switch the flow sucked into the fan
inlet between the right and left wings of a small AFC UAV (Seifert et al., 1999).
244 A. SEIFERT AND C. P. TILMANN

LIVE GRAPH
Click here to view

Fig. 11 The rolling moment generated by one wing (with a flap deflected at 20 deg)
excitation compared to the moment created by full ailerons deflection.

It was demonstrated that maximum lift and L/D were enhanced with pulsed
blowing. By sealing one of the flap-slots using adhesive tape it was possible to
generate a rolling moment that was equivalent to half the rolling moment generated
by full aileron deflection, over the entire range of relevant incidence (see Fig. 11).
The experiments of Timor et al. (2007) provide fundamental demonstrations of
roll control capabilities and 3-D aspects of separation control. These experiments
were performed on a cropped NACA 0018 airfoil, equipped with streamwise and
spanwise rows of pressure taps.
AFC excitation, emanating from 14 individually addressable piezofluidic ZMF
actuators, was applied to control the flow separating from the cutout region of a
modified NACA 0018 airfoil. The original airfoil was cropped at 70% of the origi-
nal NACA 0018 airfoil chord and at an angle of 30 deg to the local upper surface
(see Fig. 12 for a cross-section and Fig. 13 for a plan view of the airfoil, sensor,
and actuator locations). The purpose of these experiments was to study 3-D effects,
maneuverability, and redundancy aspects of segmented fluidic excitation.
Significant lift, moments (pitch and roll) and drag variations were measured as a
result of controlling the separated flow. Operating all actuators at uniform phase
and amplitude significantly increased lift and generated a pitch-down moment.
Non-uniform phase distribution along the span generated slightly higher lift
increments below stall. Significant rolling moment was generated when only half-
span of the wing was controlled. The pressure altering effect persisted up to the
LE, even though introduced close to the TE (see the images in Figs. 14 and 15,
showing surface oil flow visualization at several operating conditions and the
FIXED WING AIRFOIL APPLICATIONS 245

100

75

50

25
z [mm]

–25

–50

–75

–100
0 50 100 150 200 250 300
x [mm]
LIVE GRAPH
Click here to view
Fig. 12 A cross section of the cropped NACA 0018 airfoil, the filled circles mark the
location of the upper surface pressure taps arranged in a streamwise row; empty
symbols—lower surface. The arrow marks the 14 excitation slots located at x/c = 0.88
(Timor et al., 2007).

corresponding spanwise pressure distributions). When a pair out of the possible


14 actuators ceased operating, very little control authority was lost. This is an
important finding when redundancy of fluidic piezo actuators is considered. When
only a pair of actuators operates, a counter rotating vortex pair reattaches the flow
at the center of the controlled region and increases the severity of the separation at
its spanwise edges, in a similar manner to the effect of a mechanical VG pair.
Further upstream, the effect spreads but does not persist to the LE (Timor et al.,
2007). To understand and explain the complex 3-D flowfield, further detailed
measurements and CFD analysis are required.

E. Drag Increase—Fluidic Spoilers and Wake Manipulation


Drag increase in fully attached boundary layers, required O(1–10) control
authority, as indicated here already. If such a control authority is given, then
separating an attached boundary layer is a reasonable approach to creating an
AFC “spoiler.” These methods are complex to apply, since one must disconnect
highly interfering effects to generate pure one-axis motion. For example, simply
reattaching a separated boundary layer to one wing of an aircraft leads to both lift
increase and drag reduction, producing a roll–yaw motion. One possible steady-
state control mechanism is the jet-spoiler concept mentioned in Lachmann (1961).
Applying this mechanism is recommended at regions where the skin friction is
low and effects on the upstream boundary layer are relatively small, i.e., close to
the TE. If the boundary layer is already separated, it is possible to either excite
or enhance the shedding mechanism in the near wake. Increasing the magnitude,
shortening the formation distance or lowering the shedding frequency will
all result in drag rise. This mechanism is explained by the alteration of the
246 A. SEIFERT AND C. P. TILMANN

x/c=0.25 x/c=0.5
300 x/c=0.7 x/c=0.9
14 Actuators

LIVE GRAPH
Click here to view
200

100
z [mm]

–100

–200

–300
0 100 200 300
x [mm]
Fig. 13 Location of spanwise rows of pressure taps on the upper surface of the
cropped NACA 0018 airfoil. The dashed line marks the location of the 14 excitation
slots. Flow is from left to right. Streamwise row of pressure taps at z = 0 (Timor et al.,
2007).

vortex-induced velocity on the airfoil, altering the crossflow momentum transfer


in 2-D flows and affecting the induced drag in 3-D flows as revealed by Naim
et al. (2002, 2007) and Yom-Tov and Seifert (2005).

F. Vortex flows—Delta Wings, Fuselage, and Ahmed Car


In complex 3-D flows, streamwise vortices play an important role in the flow
evolution and momentum transfer mechanisms and therefore in their affect on the
pressure field around the body. The generation of streamwise vortices is inher-
ently a drag-producing mechanism, because of the loss of streamwise-to-rotational
momentum. However, the streamwise vortices are a primary mechanism for re-
distribution of the streamwise momentum such that the pressure field could be
affected. Due to the above-mentioned coupling and the associated complexity,
efforts to control 3-D vortex flows are quite rare. Siegel et al. (2001) and Margalit
et al. (2005) studied the application of AFC to delta wings at high incidence.
These studies concluded that vortex-breakdown was not directly affected and that
LIVE GRAPH FIXED WING AIRFOIL APPLICATIONS 247
Click here to view
–1.1 All Off
Cp x/c=0.25
All On
–1 1-7 On
–0.9 8-14On

–0.8

–0.7
x/c=0.5
–0.6

–0.5
x/c=0.7
–0.4

–0.3

–0.2

–0.1 x/c=0.9

0
–0.9 –0.6 –0.3 0 0.3 0.6 z/c 0.9

Fig. 14 Baseline and controlled spanwise pressure coefficients at a = 4 deg and


Re = 3.0 × 105. All actuators operating (squares) at Cm ª 0.02, F+ = 5 and 7 out of the
14 piezo actuators installed in the wing operate at Cm ª 0.01, note that the flow is
highly 3-D in the latter case.

Fig. 15 Surface oil-flow visualization for x/c > 0.88. Re = 0 or Re = 3.0 ¥ 105. (Same
conditions as in Fig. 14.)
248 A. SEIFERT AND C. P. TILMANN

the mechanism leading to higher lift was reattaching the separated shear-layer,
which increased the magnitude of the shear above the separated boundary layer
and its streamwise vorticity. This mechanism is not different from shortening a
2-D separation bubble.
The generic Ahmed (1983) body is an archetype of a hatchback car body. In this
complex flow, separation and roll-up of the side-wall boundary layers lead to a
dramatic drag rise at slant angles between 25 and 30 deg (see sketch in Fig. 16).
Increasing the slant angle or causing a premature boundary-layer separation off
the slant could actually lead to drag reduction due to weakening of the streamwise
vortices. These features also prevail in the flow separating from rotorcraft and
transport planes which have a relatively flat aft-loading ramp (see Fig. 17 from
Ben-Hamou et al., 2007). Reattaching the separated flow over the slant-corner
leads to lower pressure over the slant, higher rotation of the separating side-wall
boundary layers and stronger streamwise vortices that counteract the drag reduc-
ing tendency of the slant-flow separation control. In their study, Ben-Hamou et al.
(2007) combined experimental and numerical investigation to analyze the flow
around a generic transport plane/helicopter fuselage in order to reduce the drag
and alleviate unsteady loads resulting from poor aerodynamics imposed by the
presence of an aft-loading ramp. The experiment included the measurement of
surface pressures, total drag and surface oil flow, as well as off-surface smoke
flow visualization. The numerical approach applied during that limited effort
simulated only the baseline using finite volume solutions of the RANS equations,
in both steady and time-accurate modes.
The baseline flow around the fuselage model was insensitive to the Reynolds
number in the range for which it was tested. The flow separating from the aft-body
was characterized by two main sources of drag and unsteadiness. The first was a

Fig. 16 A schematic description of the flow over the aft region of the Ahmed (1983)
body. Slant angle is 30 deg.
FIXED WING AIRFOIL APPLICATIONS 249

Fig. 17 A side and top view of the simplified transport helicopter body tested by
Ben-Hamou et al. (2007). Slant angle is 30 deg (dimensions in mm).

separation bubble residing at the lower ramp corner and the second was a pair of
vortex systems developing and separating from the sides of the ramp, similar to
those of the aforementioned generic Ahmed car. The drag was elevated as the
model incidence was reduced from positive to negative angles. As the model
incidence decreases the pair of vortex systems penetrated deeper towards the
centerline of the ramp, increasing the area affected by their negative pressure.
As expected, the lower ramp corner bubble was very receptive to periodic exci-
tation introduced from the piezofluidic actuators situated at the ramp’s lower
corner, as seen in Fig. 18. Total drag was reduced by 3–11%, depending on the
model incidence, as shown in Fig. 19. There are indications that the vortex system
is tighter and the flow in the wake of the model is significantly steadier when the
bubble at the lower ramp corner is eliminated. These results, especially the control
of the bubble, should not be Reynolds number-sensitive, since the separation is
induced by surface slope discontinuity with a turbulent boundary layer upstream
of the separation region. Figures 20 and 21 show CFD and flow visualization
generated images of the vortex system, respectively.

IV. Discussion
While the previous sections provided several examples of AFC applications
representing the current state of the art, many open and fundamental issues remain.
These open issues must be resolved and an improved understanding achieved
before further progress can be made toward applying, demonstrating and fielding
AFC technology. Several open AFC issues are reviewed in the following sections.
250 A. SEIFERT AND C. P. TILMANN

LIVE GRAPH
Click here to view

Fig. 18 Baseline and controlled ramp pressures, U = 20 m/s, a = -0.5 deg, control
via four lower ramp corner actuators, total Cµ = 0.4% (Ben-Hamou et al., 2007).

Before closing this chapter on AFC applications, it is worth listing those sub-
jects that are well understood and identifying areas of boundary-layer separation
control where current understanding is lacking.

A. Resolved: Optimal Excitation Frequency


Due to the complexity of separation control in regions where transition from
laminar to turbulent boundary-layer flow is also taking place, it was initially
difficult to identify the most effective range of frequencies and propose proper
scaling. Early efforts to do so were masked by the complexity of existing actuation

LIVE GRAPH
Click here to view

Fig. 19 Drag reduction (percent of baseline drag) vs a, U = 20 m/s, lower corner


actuations (Ben-Hamou et al., 2007).
FIXED WING AIRFOIL APPLICATIONS 251

Fig. 20 Computed ramp vortex systems (Ben-Hamou et al., 2007).

systems, which were aimed at providing the required fluidic excitation but had
non-flat frequency responses. In some cases the actuation system lacked any capa-
bility at crucial frequency ranges. The nature of these actuation systems led to
identifying effective frequencies which were translated to Strouhal numbers 3–5
orders of magnitude apart (e.g., Greenblatt and Wygnanski, 2000). This situation
was altered by the series of experiments described by Seifert et al. (1996). In that
study researchers used a NACA 0015 airfoil with a simple TE flap and two excita-
tion slots (Fig. 22). The LE excitation slot was operated with the flap undeflected

Fig. 21 Baseline ramp right vortex system as seen by smoke flow visualization
(Ben-Hamou et al., 2007).
252 A. SEIFERT AND C. P. TILMANN

Fig. 22 Sketch of the modified NACA 0015 airfoil with LE (0.5 mm wide) and TE
(1.5 mm wide) slots used by Seifert et al. (1996).

at post-stall airfoil incidence, and the x/c = 0.75 slot (located directly above the flap
hinge) was operated when the flap was deflected at 20 deg and 30 deg but at low
airfoil incidence. A comprehensive calibration process was performed to result in a
flat output in terms of “dialed-in” amplitudes over a wide frequency range, compen-
sating for the dynamic response of the actuation system, ducts, and cavities in the
airfoil interior. Based on these experiments, it was clearly demonstrated that the lift
increment using low momentum coefficient, Cm = O(10-4), peaked around a Strouhal
number (termed F+, and based on the length of the controlled region, the excitation
frequency, and the free-stream velocity) of about 0.7 (see Fig. 23). Similar results
were later obtained by Greenblatt and Wygnanski (2000), see Fig. 24.

LIVE GRAPH
Click here to view

Fig. 23 The effect of low Cm (0.08%) excitation emanating from both the LE and flap
shoulder slots of the TAU-NACA 0015 airfoil over the Re range from 150,000 to
600,000.
FIXED WING AIRFOIL APPLICATIONS 253

LIVE GRAPH
Click here to view

Fig. 24 The effect of LE excitation with two amplitudes in the lift increment of the
TAU-NACA 0015 airfoil (Greenblatt and Wygnanski, 2000).

There is continued controversy about the mechanism explaining this finding,


but the effectiveness of F+  1 is most probably related to the number of vortices
residing at each instant in time above the controlled region. The minimum number
is one, with reducing efficiency when more than three vortices are present because
of their smaller size. The optimum is related to secondary effects such as boundary-
layer history, pressure gradient, curvature, free-stream turbulence and more.
Numerous experiments and simulations have demonstrated this scaling to be
correct. Moreover, an example in which F+ = 1 did not result in a positive lift
increment is yet to be found. It has also repeatedly been shown that the optimum
frequency for drag reduction is about three times higher than that for lift increment.
The reason for the difference lies in the physical mechanism enabling form-drag
reduction. For drag reduction, boundary-layer separation must be delayed to
achieve a higher static pressure and a more effective pressure recovery than that of
the baseline. This is most effectively achieved at F+ between two and three, as
clearly demonstrated by Nishri and Wygnanski (1998) and later by Naim et al.
(2007) and Pack-Melton et al. (2004, 2005, 2006).

B. Open Questions on Scaling (a partial list)


While the effects of the excitation frequency and its scaling are relatively well
understood, other important parameters such as amplitude, excitation direction
254 A. SEIFERT AND C. P. TILMANN

with respect to the crossflow, two-dimensionality vs 3-D of the excitation and the
effects of using holes vs slots for injecting the excitation are all open. In certain
cases conclusions can be drawn from past steady-state BLC studies, but the
researcher is warned about oversimplification because of the lack of dominating
flow instability in those studies. Identifying the effect of the excitation magnitude
is the most challenging open question. It is quite complex to study the amplitude
effect in isolation from other parameters because of the inherently non-linear
nature of the flow response to the excitation. It has been repeatedly found (e.g.,
Seifert et al., 1993, 1996; Nishri and Wygnanski, 1998; Greenblatt and Wygnanski,
2000) that if amplitude scans are performed at low amplitude, the universal
Strouhal law, i.e., F + = 1, is found (see Fig. 25). However, if larger amplitudes are
used, the frequency response changes. This effect was attributed by Seifert and
Pack (1999) to nonlinear excitation of a hierarchy of higher harmonics and sub-
harmonics generated by O(1) magnitude excitation. Typically, and most com-
monly, the lift is increased with increasing excitation magnitude, given that the
actuator is close to but upstream of the separation location. If this is not the case,
the required amplitude will have to overcome a threshold to become effective.
Once the flow responds to the excitation, the dCL/dCm could be O(10–100).
However, with Cm > 0.001 it is not rare to find dCL/dCm = O(1). Another limitation
is that when the separation is completely eliminated, further increase in Cm results
in lift decrement (see Fig. 26 for the ZMF case at Cm > 0.05%).
This is an indication that instability no longer plays a role in this flow condition
and that the system designers should probably employ steady-state actuators
rather than ZMF excitation (e.g, the ADVINT program, Smith et al., 2006).
Typically, when the lift changes, the moment responds in a proportional manner.
The drag, however, is more complex and initially increases rather than decreases
as Cm is increased. As indicated before, drag reduction usually requires higher
frequencies than lift recovery or enhancement, so the sensitivity to Cm is also fre-
quency dependent.

LIVE GRAPH
Click here to view

Fig. 25 The dependence of the separation (subscript s) and reattachment (subscript r)


flap deflection angles with respect to their baseline values (subscript o) vs F + at
Cm = 0.02% (Nishri and Wygnanski, 1998).
LIVE GRAPH FIXED WING AIRFOIL APPLICATIONS 255
Click here to view

Fig. 26 Lift increment vs excitation magnitude for ZMF and steady blowing (Seifert
et al., 1996).

The scaling of the excitation magnitude is another open question. While for
steady suction and vortex generator jets the velocity ratio is probably relevant, for
steady-tangential wall jets the momentum coefficient is used. This is the reason
why Seifert et al. (1993) suggested also using an oscillatory momentum coeffi-
cient for the amplitude parameter of what was later called the directed synthetic
jet by McCormick (2000). However, the scaling might be related to the vorticity
flux from the actuator (Yehoshua and Seifert, 2006a), or to a combination
__ of the
_____
velocity ratio (VR) and Strouhal number in the form H = VR /÷St = Uj /÷ fLU• as
suggested by Nagib et al. (2006), where flow instability does not play a role
because of VR ~ O(1) and higher excitation levels.
Fluidic actuators also generate vorticity flux, therefore it could be written, using
scaling arguments, that a vorticity flux coefficient should be defined in the form

È dUp (t , y) ˘
ÍUp (t , y) ˙ 2 2
Î dy ˚ max U p2 q 2q Ê Up ˆ 0.01c Ê Up ˆ
CVF ∫ ª = ª (1)
È dUBL ( y) ˘ h 2 U e2 h ÁË Ue ˜¯ h ÁË Ue ˜¯
U
Í BL ( y ) ˙
Î dy ˚ max

In the above, Up is the slot exit velocity and its wall-normal derivative and UBL
is the boundary-layer velocity and its wall-normal derivative. Using the scaling
arguments above, it is found that for a given slot width, h, boundary-layer
256 A. SEIFERT AND C. P. TILMANN

momentum thickness is q, Up and Ue (external to the BL free-stream velocity).


Using the accepted definition of the momentum coefficient
2
2h Ê Up ˆ
Cm ∫ (2)
c ÁË Ue ˜¯

and the values of: h/c = 0.005 (and q/c = 0.005), one may conclude that the ratio
of the vorticity flux to momentum coefficient is of the order

0.01c 2h 0.005c 2 5 ¥ 10 -3
CVF Cm ∫ = ª = 200 (3)
h c h2 2.5 ¥ 10 -5

This ratio certainly requires some consideration as to which excitation magnitude


parameter one should refer to.
Further study, especially with different slot widths, will be required to settle
this issue. Certainly the effects of the 3-D distribution of amplitude, phase, and
frequency are open, and their effects have been discussed throughout this
chapter.

C. Open Issues in System Studies


The primary weakness of any system study lies in the assumptions it uses in
order to be able to predict performance and trade-offs based on an incomplete
knowledge-base as to how the system would function. A dangerous path followed
in many system studies is to linearly extrapolate experimental results obtained at
small scale and off-design conditions. This can lead to confusing and often dis-
appointing results as the system is matured toward full-scale demonstration.
A primary shortcoming to applying AFC is that it has been most commonly
considered as a fix and retrofit rather than an essential toolbox in the preliminary
design stages. The design approaches of Glauert (1947), Glauert et al. (1948),
Goldschmied (1981, 1987), Leibeck and Ormsbee (1970), and Stratford (1956)
have not yet been suitably extended to include AFC techniques in designing new
aerodynamic shapes that will take advantage of the proven capability of AFC to
reattach separated flow or delay the separation of an attached boundary layer.

V. Summary and Recommendations


The focus of this chapter has been to introduce the reader to several applica-
tions of active boundary-layer separation control. A brief introduction led to a
review of the pioneering system study conducted by Boeing in 1999 and to the
description of several AFC demonstrations that followed. Through these studies it
has been shown that the excitation Strouhal number should be of order unity,
based on the imposed frequency and the length of the separated region. The
Reynolds number is not a limiting factor; probably the opposite is true. It requires
lower excitation magnitude to control boundary-layer separation as Re increases.
Turbulence also does not really hinder success when the scaling parameters are
understood. Again, scaling is simpler when boundary-layer transition is not a
FIXED WING AIRFOIL APPLICATIONS 257

factor. Many issues remain open, and to enable progress we recommend the
following.
Publish failures: Although it is not the purpose of this chapter, “failed” studies
should be fully analyzed and, more importantly, reported. In using the quotation
marks above, it is meant that in many studies the expected results were not
obtained. However, practitioners in the field could learn a great deal from such
expertly conducted studies that did not result in the hoped for outcome.
Conduct multidisciplinary collaborations: The complex flow physics of CLAFC
systems requires the incorporation of knowledge from multiple fields. It is clear
that fruitful interaction between different groups of scientists, along with patient
and guiding industry collaboration and support from the private sector—not only
government—is a crucial enabler.
Conduct comparative studies: To enable the enormous progress that is required,
the role of government-funded research is to push the limit of knowledge and, at
the same time to enhance the effectiveness and foster the application of already
existing research. This is best done by open calls for comparative studies, promot-
ing the research of several key problems in a parallel study, but with different tools
and conducted by different groups. The results of such studies—such as the recent
Unsteady Flow Control CFD validation workshop (Rumsey et al., 2006)—will
significantly promote the state of the art, and will also reveal the shortcomings of
several popular, perhaps even trendy, approaches.
AFC will be beneficial when effectively coupled with flow instability in one of
two possibilities. The simpler possibility is as a retrofit or a design of a system
where considerations other than efficiency have dictated the shape, e.g., stealth,
design of short and tall cars, SUVs and truck-trailers. The more complex path—
but one possibly leading to a breakthrough in performance—is incorporation in
the design process of new shapes that will be optimized for the use of unsteady
boundary-layer separation control, distributed loading, and highly integrated
(probably even distributed) propulsion.
Chapter 9

Turbomachinery Applications

Hermann F. Fasel* and Andreas Gross†


University of Arizona, Tucson, Arizona
Jeffrey P. Bons‡
Ohio State University, Columbus, Ohio
and
Richard B. Rivir¶ and Rolf Sondergaard§
Air Force Research Laboratory, Wright–Patterson Air Force Base
Dayton, Ohio

I. Introduction and Motivation


In the everlasting quest for greater efficiency of jet engines, modern low-pres-
sure turbines (LPTs) have to drive larger fans at lower speed. At the same time,
engine complexity and engine weight need to be reduced. One way to reduce
engine weight is to lower the stage solidity (inverse of the blade spacing) without
compromising performance. LPTs must work efficiently over a large range of
chord Reynolds numbers. Typically, the Reynolds number is larger during take-
off, where the engine is running at full power and the air density is high. During
cruise at higher altitudes, the air density is low and the engine is running at a lower
cruise power setting. Compared to take-off, the LPT is operating at lower Reynolds
numbers. For low Reynolds number conditions or when more aggressive designs
are considered, laminar separation can occur on the suction side of the blade,
resulting in significant turbine and overall engine performance losses. Separation
occurs when the flow cannot negotiate the adverse pressure gradient associated

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Professor, Aerospace and Mechanical Engineering Department. Member AIAA.
†Assistant Research Professor, Aerospace and Mechanical Engineering Department. Member

AIAA.
‡Professor, Aerospace Engineering Department. Associate Fellow AIAA.
¶Senior Scientist, Propulsion Directorate. Associate Fellow AIAA.
§Senior Research Engineer, Turbine Branch. Member AIAA.

259
260 H. F. FASEL ET AL.

Fig. 1 Laminar separation in aft part of LPT blade. Iso-contours of spanwise vorticity
(Reprint from Gross and Fasel, 2005b).

with the pressure recovery in the aft part of the blade. Separation is often said to
be a result of “uncovered turning” (Fig. 1).
In fact, Sharma (1998) reported a near 300% increase in the wake pressure loss
coefficient at Reynolds numbers below 95,000 compared to the loss coefficient at
higher Reynolds numbers. This increase was found to be primarily due to separa-
tion occurring over the trailing 50% of the suction surface. The wake pressure loss
coefficient

p0,in - p0,out
g = (1)
1
ruin2
2

is defined
____
as the difference between the stagnation pressure at_
the inlet and outlet,
p0,in - p0,out, normalized by the inlet dynamic pressure, 0.5ru2in. The overbar indi-
cates an average over the cascade plane. The wake pressure loss coefficient is a
measure of the total pressure loss incurred by the flow as it turns through the blade
row in the frame of reference of the blade row. The higher the number, the larger
the losses incurred. The adverse pressure gradient depends on the aerodynamic
loading of the blade and is larger for more aggressive designs. Large adverse
pressure gradients can not only lead to separation but also “accelerate” transition
from laminar to turbulent flow. Transition, in turn, can prevent separation because
of the larger wall normal momentum exchange of turbulent boundary layers when
compared with laminar boundary layers. Earlier transition may prevent or reduce
separation but it will also result in larger wall heat loads and friction losses. At low
Reynolds number conditions, the laminar boundary layer is hydrodynamically
more stable, resulting in later transition which, as a consequence, favors separa-
tion. Free-stream turbulence (FST), surface roughness, and unsteady wakes (shed
by upstream stages) among others can accelerate transition by “by-passing”
(Morkovin, 1969) the linear transition stages. In case of laminar separation, as a
result of a shear-layer instability, the separated laminar boundary layer becomes
unsteady and may transition to turbulence. The shear-layer instability ampli-
fies 2-D disturbances, resulting in a “roll-up” of the shear-layer into spanwise
TURBOMACHINERY APPLICATIONS 261

coherent vortical structures. Secondary instability mechanisms lead to amplified


3-D disturbances which may ultimately result in breakdown to turbulence, and
thus lead to a weakening of the spanwise coherent structures. The extent of the
transition region is larger at low Reynolds number conditions and affected by
other external factors such as FST and passing wakes. Dependent on the severity of
the separation and the promptness of the transition process, the flow may reattach
(enclosing a so-called laminar separation bubble) or not reattach, resulting in
complete stall and severe performance losses. Among many other additional
effects that were disregarded in this brief discussion but need to be considered
when implementing flow control into turbine blades are the rotating flow effect,
the three-dimensionality of the blade geometry (turbine blades have a small aspect
ratio) and the tip gap leakage flow (resulting from the gap between the tip of the
blade and the housing). All of these affect the fluid dynamics and especially tran-
sition and separation. In summary, the various mechanisms, in particular the deli-
cate interaction of separation and transition, that govern the flow dynamics and
hence the performance of the blade are complicated and intertwined with each
other. However, many aspects of the problem and the sensitivity of the flow to
external disturbances also make it accessible for active flow control (AFC). The
experimental and numerical studies which are reported in the remaining part of
this chapter do not consider 3-D effects. This simplifies the setup of the experi-
ments and simulations as well as the analysis. It is implicitly assumed that the
physical mechanisms responsible for the success of an AFC strategy are qualita-
tively similar when 3-D effects are considered.
The intensity and spectrum of FST which has its origin in the combustor sec-
tion of the engine are in most instances unknown. Free-stream turbulence can
result in an earlier transition of the flow through a classical transition scenario or
for large FST intensity (FSTI) through a “by-pass” transition process (Morkovin,
1969) where turbulent spots appear. Schubauer and Klebanoff (1956) described a
calmed, quiet, low-loss region which characteristically follows behind a turbulent
spot. The first in-detail measurements focusing on transition and separation in
LPTs as well as the definition of the turbine boundary-layer quiet region were
provided by Halstead et al. (1995). The calmed regions, which appeared to follow
the turbulent spots produced in the wake’s paths, were found to be effective in
suppressing flow separation. Later, in experiments carried out at the Von Kármán
Institute (VKI) the transition delay caused by the quiet region was exploited in an
LPT cascade where every seventh blade was removed and which was run at full
Reynolds and Mach number without flow separation. A number of related publi-
cations by Halstead et al. (1997a–e) dealing with other aspects of LPT aerody-
namics followed. Halstead and colleagues found peak turbulence intensities of
about 5% in the rotor wakes and 3.5% in the wakes from the first stator (measured
at the second stator of a three-stage LPT). Between the upstream stator and rotor
wakes, the minimum level of turbulence intensity can be as low as 1.5%, depend-
ing on the circumferential position of the sensor. Halstead et al. (1997a–e) also
observed that, in general, turbulence intensities increase throughout the turbine.
The increase is mainly attributed to the “filling in” of the free-stream between the
wakes of the preceding blade rows by the effects of the more upstream wakes.
However, the wakes of the nearest upstream blade row tend to remain dominant
with regard to the effect on the blade-surface boundary layers.
262 H. F. FASEL ET AL.

Corrosion, deposition, pitting, or erosion, and de-bonding or spallation (of


thermal barrier coatings) which are most prominent immediately downstream of
the combustor, where highly turbulent combustor exit flows spew hot combustion
products and other airborne particulates at the turbine surfaces, can result in an
order of magnitude increase in surface roughness. A systematic catalogue of the
different types of surface roughness typically encountered in turbine applications
was compiled by Bons et al. (2001b). Depending on the operating conditions,
roughness can both deteriorate efficiency (typically in high Reynolds number
attached flow conditions by increasing drag) or improve performance (in low
Reynolds number conditions by reducing flow separation). Turbine blades operate
in the turbulent wake of upstream stators (or vanes). Experimental investigations
of the wake influence on boundary-layer development for high-lift LPT blades
(e.g., Reimann et al., 2007; Wolff et al., 2000) showed that the transition region is
shifted upstream and that laminar separation is diminished or suppressed by
unsteady wakes. The region of attached turbulent flow is followed by a calmed
flow region that is nonturbulent and very resistant to separation. The wake passing
frequency was found to have a significant influence on boundary-layer develop-
ment. In their review paper, Hodson and Howell (2005) provide a detailed discus-
sion of wake-induced transition and its effect on separation. This effect has already
been exploited for ultra-high-lift profiles (e.g., Howell et al., 2002; Zhang and
Hodson, 2005).
At low Reynolds number conditions the wake may not be strong enough to
completely suppress LPT separation. In these instances, additional performance
improvements may be obtained with AFC. Two possible scenarios can be envi-
sioned. First, AFC could be employed in an on-demand fashion to assist during
critical off-design conditions and then be deactivated during nominal operation.
This strategy, however, is less relevant for commercial jet engines where the low
Reynolds number conditions occur during cruise which is the most important
design point. A different strategy would be to reduce the blade count (and thus
component weight) and avoid separation at all flight conditions, particularly dur-
ing cruise, by continuously applying AFC. When AFC is implemented success-
fully into an existing (conventional) blade design, the stage solidity can be reduced
without compromising overall performance (Sondergaard et al., 2002b). Even
larger gains may become possible when the integration of AFC for LPT blades is
considered from the very beginning of engine development, and not just as an
add-on or fix (e.g., Bons et al., 2005). In summary, AFC when applied to LPT
blades may ultimately result in true breakthroughs in the efficiency and overall
performance of jet engines.
The remaining part of this chapter is structured into three parts. Section II pro-
vides a summary of LPT experiments at the Air Force Research Lab (AFRL) at
Wright Patterson Air Force Base and at Brigham Young University. In these
experiments particular emphasis was put on separation control by vortex genera-
tor jets (VGJs). Section III reports on numerical simulations conducted at the
University of Arizona. These simulations provide detailed insight into the complex
fluid dynamics and flow physics that cannot be obtained from the experiments
alone but only in combination with high-fidelity numerical simulations as dis-
cussed in this chapter. The chapter concludes with a discussion of the results, both
experimental and numerical, in Sec. IV.
TURBOMACHINERY APPLICATIONS 263

II. Experimental Investigations


A. Overview
The reduction in separation bubble size and hence profile losses for increased
levels of FSTI was demonstrated in various LPT experiments (e.g., Huang et al.,
2006a; Sondergaard et al., 2002a). Separation can also be delayed by carefully
placed roughness elements which “trip” the boundary layer to turbulence, result-
ing in an increased turbine blade loading (Huang et al., 2006a; Lake et al., 1999).
However, increased viscous losses at conditions where unmodified blades yield
satisfactory turbine performance are a disadvantage of such passive techniques. In
an extensive experimental research program at the AFRL at Wright-Patterson
AFB, Rivir and co-workers (Bons et al., 1999, 2001a, 2002; Sondergaard et al.,
2002a, b) systematically investigated the benefits of AFC with steady and pulsed
VGJs for a linear Pack B LPT cascade. The Pack B profile is a low Mach number
scaled high-performance blade profile originally released by Pratt and Whitney.
For VGJ control, fluid is injected into the boundary layer through a spanwise array
of small holes placed along the spanwise direction on the suction side of the blade,
upstream of the separation line. Pulsed VGJs were found to be considerably more
effective than steady VGJs. Steady blowing was shown to generate streamwise
vortices (similar to conventional vortex generators) leading to free-stream momen-
tum entrainment (Eldredge and Bons, 2004), whereas pulsed blowing was shown
to cause early boundary-layer transition, especially when the jets were employed
near the “natural” (uncontrolled) separation location (Bons et al., 2001a). The
wake loss coefficient could be reduced by more than 50% using pulsed VGJs
(Bons et al., 2001a). Control remained successful even when the blade spacing
was increased by 50% (Sondergaard et al., 2002b). In related research at Brigham
Young University, Bons et al. (2005) showed that more aggressive LPT blade
designs were possible when considering AFC from the very beginning. The
complementary roles of transition and streamwise vorticity in aiding the separa-
tion control phenomenon have also been explored in detail (Reimann et al., 2006,
2007). Finally, the influence of upstream wakes was explored and criteria were
proposed for the optimum integration of pulsed flow control into the embedded
turbine row flow environment (Bloxham et al., 2007). Extensive experimental
studies of active LPT separation control using steady and pulsed plasma actuators
were conducted by Corke and co-workers (Corke and Post, 2005; Huang et al.,
2006a, b). In these experiments a drastic reduction in boundary-layer separation
on the suction side of the LPT blade was achieved. Pulsed actuation was shown to
be much more efficient than steady actuation and to require a lower energy input.
A modulation frequency that generated approximately two spanwise vortices over
the length of the separation zone was found to be optimal.

B. Low-Pressure Turbine Separation Control using Passive Actuators


Many passive separation control techniques involve the introduction of longitu-
dinal or streamwise vortices in some form and their subsequent interactions in
boundary layers. Studies of such techniques go back over 50 years (Arts and
Colton, 2004). The most well known passive vortex generation techniques include
half delta wings or fences which are currently employed extensively on external
264 H. F. FASEL ET AL.

aircraft flows. Riblets, studied by Walsh (1990), reduced drag by typically 6–8%.
Bearman and Harvey (1976, 1993) employed dimples on golf balls to reattach
separated flows and reduce drag. Mahmood et al. (2000) and Musiyenko (1993)
applied dimples to flat plate boundary layers to increase heat transfer in wall-
bounded flows. Large-eddy breakup devices, roughness, and turbulent trips were
used by Musiyenko (1993), Walsh (1990), and others to reattach separated flows
and increase the critical angle of attack. The most heavily studied passive LPT
flow control techniques are the clocking of the rotor stator wakes and the applica-
tion of dimples.
Clocking or indexing relates to the relative circumferential position of fixed and
rotating blade rows in consecutive stages, the clocking position. Clocking influ-
ences the interaction of the wakes from upstream blade rows with the boundary
layers on downstream blade rows and, thereby, performance. In particular, both
experimental (e.g., Huber et al., 1996; Sharma and Tanrikut, 1994) and numerical
(e.g. Cizmas and Dorney, 1998; Dorney and Sharma, 1996; Dorney et al. 1998;
Griffin et al., 1996) investigations showed how the relative position of the rows,
together with the blade count ratio between consecutive rows, influence the
boundary layers on the downstream rows. In particular, wakes from upstream
stages were found to reduce flow separation and therefore increase performance.
It was demonstrated that larger efficiency benefits can be achieved if the blade
ratio between consecutive blade rows is near 1 : 1, while practically no effect can
be detected if it is far from unity. Dimples reduce separation by generating longi-
tudinal vorticity which reenergizes the boundary layer by entraining higher energy
boundary-layer and free-stream fluid. Lake et al. (1999) applied dimples to a Pack
B cascade to eliminate separation at low Reynolds numbers with smaller losses
than incurred with trips.
The LPT is responsible for driving the large fans that can produce up to 80% of
the total engine thrust in high bypass ratio engines. Since modern LPT efficiencies
are already very high (over 90%) further increases are difficult. However, a reduc-
tion of the LPT weight, which typically accounts for about 30% of the overall
engine weight, promises further significant savings. The LPT weight can be low-
ered by reducing the number of blades needed for a given work extraction, which
increases the blade loading. Increased loading, however, can lead to increased
aerodynamic losses due to strengthened end-wall flows and boundary-layer sepa-
ration. These phenomena are amenable to flow control. Reductions of 12% in
blade count were shown by Howell et al. (2002) and increases of 20% in blade lift
without an increase in blade profile loss were obtained by Curtis et al. (1997) by
adjusting the clocking such that upstream wakes were exploited for separation
control of highly loaded airfoils. In the presence of more severe flow separation,
application of dimples slightly upstream of the separation location has been shown
to be effective in reattaching the flow and preventing separation on highly loaded
airfoils.
The main losses in the LPT occur in the profile of the boundary layer. In the
early 1990s, Halstead et al. (1995) (see also Halstead’s PhD dissertation, 1996)
and Hodson et al. (1993) with his colleagues at Cambridge and Rolls Royce stud-
ied the development of axial turbine boundary layers with the goal of reducing the
boundary-layer losses and preventing separation of highly loaded airfoils. It was
found that for highly loaded airfoils separation was followed shortly by transition
TURBOMACHINERY APPLICATIONS 265

caused by the passing wakes. In particular, the wakes were found to generate tur-
bulent spots with leading and trailing edges that were traveling at roughly 50%
and 30%, respectively, of the free-stream velocity. A calmed region with a full
velocity profile was observed behind the turbulent spot that was very resistant to
separation. Appropriate clocking of the stator wakes, taking advantage of transi-
tion and the reduced losses of the calmed region, allowed researchers in one
instance to increase the blade loading by removing 20% of the LPT blades while
still maintaining fully attached flow (Curtis et al., 1997).
Passive flow control (PFC) techniques have similar frequency and coupling
requirements to AFC approaches. Shedding frequencies and characteristic dimen-
sions, such as the separation bubble length, are still important but are coupled in a
natural straightforward way by being a part of the wall process. Yurchenko and
Rivir (2000), when studying the stability of boundary-layer flows and Görtler
vortices, found that stable longitudinal vortices could be introduced into the flow
in the near wall region if a frequency or a scale smaller than the fundamental most
unstable frequency using vortex generators was chosen. Using a compressible
2-D Navier–Stokes solver, Wu et al. (1998) showed up to 70% increases in
unsteady airfoil lift for effective forcing frequencies of 0.3–2 times the shedding
frequency at Reynolds numbers of 5 × 105. The parameter range for stable, “long-
living” vortices (Rivir et al., 2004) is quite large and compatible with the favorable
interaction frequencies or scales of the DNS by Wu et al. (1998). A spanwise
periodic pressure distribution is common to both the calculations and experiments
employing dimples and riblets.
Dimples provide an effective method for reenergizing a separating boundary
layer. For example, dimples applied to golf balls reduce separation at low Reynolds
numbers resulting in significant drag savings. Dimples shed multiple vortices
as illustrated in Fig. 2. The basic concept is to introduce longitudinal vortices into
the boundary layer. Depending on the pressure gradient, streamwise vortices can
become stretched in the downstream direction and remain confined near the wall
(Fig. 2). Streamwise vortices prevent flow separation by entraining high momentum
free-stream fluid into the boundary layer. Lake (1999) and Lake et al. (1999) reported
successful reattachment of separated low Reynolds number flows with dimples.

Fig. 2 Elliptical dimple shedding multiple pairs of counter-rotating longitudinal


vortices (Reprint from Mahmood et al., 2000).
266 LIVE GRAPH H. F. FASEL ET AL.
Click here to view

Fig. 3 Comparison of dimples and spanwise V groves with clean blades; FSTI = 1%.

They compared dimples located at 50, 55, and 60% chord with V groves, sand
grain turbulent trips, and wire trips. Fig. 3 shows that a 60% reduction of the loss
coefficient can be obtained with dimples at a Reynolds number of 45,000. These
measurements were subsequently carried out at Reynolds numbers down to
25,000. Dimples located at 55% chord, which was just ahead of the separation
location at 63% chord, performed best.
Using the multiblock CFD code Swift, Cizmas and Miller (2000), modeled the
experiment with and without dimples and found that dimples reattached the flow
despite the massive separation without dimples. Rouser (2002) computed Lake’s
elliptical dimples using the commercial CFD code Fluent. These calculations were
used to investigate the effects of the dimple shape and spacing. Rouser’s computa-
tions indicated that wider dimple spacing would still be effective and that further
changes in dimple shape did little to enhance the dimple performance (Casey,
2004). Multiple rows showed no additional improvement in cascade performance.
Spacing, shape, and configuration trends were all confirmed by experiments.
Passive control of LPT flows has been shown to be an effective technique for
reducing LPT weight without penalizing turbine performance. The allowable
design space for effective passive LPT flow control is large and supported by both
experiment and theory. Improved understanding of the development of the turbine
boundary layers, transition, and turbine turbulence have been important in achiev-
ing useful PFC techniques for LPT flows.

C. LPT Separation Control using Steady Jet Actuation


Boundary-layer separation control in diffusing flows (under pressure conditions
similar to the aft portion of a turbine blade) has been studied in the laboratory for
TURBOMACHINERY APPLICATIONS 267

Fig. 4 Schematic of vortex generator jets with skew angle q and pitch angle, j. VGJs
are typically placed along the span with a constant spanwise spacing, Dz (Reprint
from Gross and Fasel 2005a).

many years. Lin et al. (1990) presented results from a number of passive and active
strategies employed in a backward-facing, curved-ramp wind tunnel facility.
Strategies investigated include submerged vortex generators, large-eddy breakup
devices, elongated boundary-layer arches, flush-mounted Helmholtz resonators,
and VGJs. Of those studied, only the last two, Helmholtz resonators and VGJs, are
active techniques. These have the advantage that they can be turned off when not
required for flow control. This is especially desirable for turbine blade applica-
tions, since any passive control strategy which is successful at low Reynolds
numbers would be likely to increase the blade’s drag penalty and surface thermal
loading at higher (non-separating) Reynolds numbers. Of the active strategies,
only VGJs had a significant effect on reducing diffuser separation in the study by
Lin et al. (1990).
Steady VGJs were also used extensively in separation control experiments on a
flat plate under an adverse pressure gradient by Johnston and Nishi (1990) and
Compton and Johnston (1992). Their work demonstrated that VGJs are particu-
larly well-suited for boundary-layer separation control. The blowing ratio B is a
measure of the intensity of the VGJs and defined as the ratio of the maximum jet
exit velocity, vjet, to a reference velocity, vref,*

vjet
B= (2)
vref

VGJs are typically configured with a low pitch angle j (30–45°) and aggressive
skew angle q (45–90°) to the free-stream flow direction (Fig. 4). Here the pitch
angle is defined as the angle the jet makes with the local surface and the skew
angle is defined as the angle of the projection of the jet on the surface relative to
the local free-stream direction at the point of injection. Depending on the blowing
ratio B in this skew configuration, each VGJ creates a horseshoe vortex pair with
one very strong leg accompanied by a weak leg of opposite sign. The result is a
single, dominant streamwise vortex which slowly decays in the downstream
direction, rather than the two relatively weak counter-rotating horseshoe vortices
generated by a jet with zero degrees skew angle or a passive boundary-layer

*In experiments the reference velocity was taken as the local free-stream velocity. For the numeri-

cal results, the cascade inlet velocity vin was chosen as reference velocity.
268 LIVE GRAPH H. F. FASEL ET AL.
Click here to view

Fig. 5 Streamwise wake velocity profiles for steady VGJs normalized by mid-channel
velocity. Traverses across exit wakes at 0.62 axial chords downstream from blade trail-
ing edges. Re = 42,000; FSTI = 1%; B = 0, 1, 2, & 4 (Reprint from Bons et al., 1999).

obstruction (vortex generator). Johnston and Nishi (1990) showed that in a 2-D
planar diffuser, this single-sign vortex energizes the separating boundary layer by
effectively bringing high momentum free-stream fluid down to the wall. The suc-
cess of this control strategy has been demonstrated with both laminar and turbulent
boundary layers (Lin et al., 1990), although the vortex structure is subtly different
for the two cases.
The application of VGJs to LPT separation control was first demonstrated by
Bons et al. (1999) in a low-speed linear cascade using the Pack B profile. The
eight-blade cascade was fitted with steady VGJ actuators on the fifth blade only.
Figure 5 shows the normalized velocity wake profiles for blade 5, the controlled
blade, and blade 6, an unmodified blade, for blowing ratios (B) of 0, 1, 2, and 4.
The vortex generator jets completely eliminated the suction side separation zone
leaving only a narrow wake. A graphic illustration of the bulk flow adjustment
caused by the VGJs was obtained using smoke flow visualization captured with a
CCD camera. Figure 6 shows two images taken of a smokeline visualization of
the flow near the suction surface of the controlled blade at Re = 42,000 for two
blowing ratios (B = 0 and B = 2). The use of VGJ control brought the smoke-
tagged streamline closer (and nearly parallel) to the surface of the blade and
reduced its downstream mixing.
The flow visualizations in Fig. 7 demonstrate that the flow can also be success-
fully reattached with steady blowing at the lower Reynolds number of 25,000. The
flow visualization for the uncontrolled case shows the separated flow region and a
roll-up of the separated boundary layer into spanwise vortices. For blowing with
B = 1 regularly spaced structures are visible. For B = 2 more random structures
are seen. The flow response is clearly different depending on the blowing ratio.
TURBOMACHINERY APPLICATIONS 269

Fig. 6 CCD images of smoke line flow visualization between VGJ blade (5) and
uncontrolled blade (6). Flow is right to left. Re = 42,000; FSTI = 1%. a) No blowing,
B = 0 and b) blowing, B = 2 (steady VGJs).

The flow structures and their dynamics will be discussed in more detail in Sec. III
(simulations).
The wake velocity profile in Fig. 5 was integrated across the controlled blade
wake to arrive at a single quantitative measure of the VGJ effectiveness, the wake
momentum deficit. This wake loss parameter is plotted vs VGJ blowing ratio in
Fig. 8. The loss parameter is normalized by its uncontrolled value (B = 0) such
that values less than unity indicate effective wake control. For these test condi-
tions, a minimum steady blowing ratio of 1.5 yielded the maximum separation
control of 60% reduction in wake momentum deficit. Also shown in Fig. 8 are
data from Eldredge and Bons (2004) taken in a larger-scale three-blade linear
cascade facility constructed at Brigham Young University for detailed boundary-
layer velocity measurements of the controlled flowfield. Figure 9 shows bound-
ary-layer data obtained in this facility superposed on the Pack B blade profile.
The controlled boundary layer (B = 4) is shown to be attached down to the last

Fig. 7 Laser sheet smoke flow visualizations. Flow is left to right. Re = 25,000;
FSTI = 1%, steady VGJs.
270 H. F. FASEL ET AL.
LIVE GRAPH
Click here to view
1.2

1
Normalized Wake Deficit

Bons et al., steady VGJ @63%Cx


0.8 BYU, steady VGJ @59%Cx

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3 3.5 4
B

Fig. 8 Normalized wake deficit vs blowing ratio (B) for steady VGJs. BYU data
from Eldredge and Bons (2004) compared to Bons et al. (1999).

68% cx

VGJs

77% cx

82% cx

B= 0
ooo B= 4
87% cx
-..- Attached boundary layer
- - - Separated boundary layer

Fig. 9 Mean boundary-layer velocity profiles for steady VGJs at B = 0 and 4 super-
posed on the Pack B blade profile. Flow is left to right. Re = 25,000; FSTI = 1%.
TURBOMACHINERY APPLICATIONS 271

Fig. 10 Streamwise velocity (u/Uin) contour plots taken at 68, 77, 82, & 87% Cx for
steady VGJs. VGJ locations at 63% Cx indicated by arrows at left. Jet-induced stream-
wise vortex migration indicated with dot and curved arrow. B = 4, Re = 23,500.

measurement station (87% Cx). Velocity maps (Fig. 10) from the last four mea-
surement stations in Fig. 9 show the locations of low-momentum jet fluid as well
as the streamwise vortex core (indicated with black dot and curved arrow) as
measured using two-component hot wire anemometry. The vortex migrates span-
wise in the direction of injection and is coherent clear to the trailing edge (TE) of
the blade.
Sondergaard et al. (2002a) expanded the application of VGJs to the Pack B
over a broader range of injection locations (45–83% Cx). They found VGJs to
be effective over the full range of locations except for 83% Cx. This position
was considered to be too far downstream of the natural separation location
(70% Cx). Sondergaard and colleagues also studied elevated FST levels up to
4% and found steady VGJs to be effective, though at a reduced level due to the
smaller separation zone at higher FST levels. Olson et al. (2005) increased the
FSTI to 10% and found minimal VGJ effectiveness due to the smaller separa-
tion, even at low Reynolds numbers. The smaller separation zone was caused
by premature boundary-layer transition induced by the elevated turbulence in
the free-stream.

D. Low-Pressure Turbine Separation Control using Pulsed Jet Actuation


Given the obvious engine cycle costs of implementing VGJs in a jet engine,
pulsed VGJs offer the possibility of comparable control effectiveness at greatly
reduced required massflow. Also, the prospect of synchronizing pulsed jets with
the blade passing frequency may lead to synergies between the application of
control and the separation zone’s response to the convected wake disturbance
from upstream blade rows. As such, an initial study was conducted by Bons et al.
(2001a) to explore the effect of unsteady VGJs at a dimensionless forcing
272 H. F. FASEL ET AL.
LIVE GRAPH
Click here to view
Normalized Integrated Wake Loss Coefficient 1.2

Steady @63%Cx
0.8
Pulsed @63%Cx

0.6

0.4

0.2

0
0 1 2 3 4
Mean Blowing Ratio (B)

Fig. 11 Integrated wake loss coefficient (g int) normalized by loss coefficient for B = 0
vs mean blowing ratio (B). Data for steady vs pulsed blowing at 10 Hz and 50%
duty cycle (Reprint from Bons et al., 2001a).

frequency (F+) of 0.3. As shown in Fig. 11, pulsed blowing has a much lower mini-
mum blowing ratio for effective separation control compared to steady blowing.
The data in Fig. 11 is for actuation at 63% Cx, but similar results were found for
pulsed actuation further upstream at 45% Cx as well.
Whereas the relevant physical mechanism for steady blowing effectiveness is
clearly the generation of strong coherent streamwise vortices (Fig. 10) that
entrain high momentum free-stream fluid and reenergize the separated bound-
ary layer, the mechanism responsible for pulsed control is less obvious. The
result in Fig. 11 suggests that a disturbance of any magnitude (i.e., mean blow-
ing ratio) produces an effective control by exploiting a hydrodynamic instability.
Thus, the exact amount of fluid injected with the pulse may be of secondary
importance to the destabilization of the boundary layer. To explore this hypoth-
esis, Bons et al. (2001a) conducted a second set of experiments with a variable
jet duty cycle. The duty cycle time signal, f(t), is one during a fraction t of the
period T, which is the inverse of the pulsation frequency, T = 1/F+ (Fig. 12).
The frequency, F+, is non-dimensionalized with scales that are relevant from the
hydrodynamic stability point of view, such as the length of the separated flow
region and the free-stream velocity.*

*In the experiments the forcing frequency was non-dimensionalized by an averaged free-stream

velocity (averaged in streamwise direction between VGJ location and TE) and the distance from the
VGJ location to the TE. In the simulations the forcing frequency was non-dimensionalized with inlet
velocity and axial chord length.
TURBOMACHINERY APPLICATIONS 273

Fig. 12 Reduced duty cycle time signal.

The jet exit velocity is vjet = Bvref f(t) where the blowing ratio B is defined as the
ratio of maximum jet exit velocity to reference velocity. In the experiments blow-
ing ratios in the range of 1–4 and duty cycles in the range of 1–50% were explored.
The momentum coefficient

Ê1 ˆ
Ú ÁË T Ú v dt˜¯ dA
2
(3)
cm =
v Cx D z
2
in

is commonly used to describe the energy expense of the forcing. It compares the
time-averaged momentum of the actuation integrated over the area of the actuator
A to the inlet momentum, v2in, times a reference area, Cx Dz.
The duty cycle time signal can be Fourier decomposed

a0 È nt ˘ (4)
f (t ) =
2
+ Âa
n
n cos Í2p ˙
Î T˚

with Fourier mode amplitudes a0 = 2t, an = 2/p n sin p nt, and frequencies fn = n/T.
As the duty cycle approaches very small values, the amplitude drop-off towards
higher frequencies becomes smaller, meaning that for smaller duty cycles higher
harmonics are forced at comparatively larger amplitudes. Also noteworthy is that
for a duty cycle of t = 0.5 only uneven multiples of F+ = 1/T are forced. Arguably,
a perfect square forcing function cannot be achieved in the experiment. Figure 13a
shows jet exit velocity histories for duty cycles from 1 to 100%. Figure 13b shows
the corresponding effectiveness. From this data it appears that the beginning and
end of the jet pulse (and not the injected mass itself) provide the means for influ-
encing the boundary layer. Bons et al. (2001a) proposed that this is done through
the mechanism of early boundary-layer transition and that the vortical entrain-
ment of free-stream fluid only played a secondary role.
A follow-on study by Bons et al. (2002) explored the role of frequency and
found VGJs to be effective over nearly two orders of magnitude in reduced fre-
quency (0.1 < F+ < 10). More importantly, Bons and colleagues identified a phase
lag, or relaxation, in the response of the boundary layer to the VGJ-induced
reattachment of the separation zone. They postulated that the sequence of events
affecting the success of pulsed VGJ control includes an initial transient as vortic-
ity is shed due to the increased blade circulation. This is followed by the VGJ
“on” (or controlled wake) cycle and then a final relaxation period. The relaxation
period is related to a physical time scale of the flowfield and acts as an effective
274 H. F. FASEL ET AL.
LIVE GRAPH LIVE GRAPH
a) Click here to view b) Click here to view
1.2

Normalized Integrated Wake Loss Coefficient


2.5

1
Instantaneous Jet Blowing Ratio

100% (steady) 0.8


1.5 50%
25%
0.6
10%
1 5%
1% 0.4

0.5 0.2

0
0
0 0.2 0.4 0.6 0.8 1 0 20 40 60 80 100
Forcing Duty Cycle [%]
Time / Forcing Period
Jet exit velocity history Loss coefficient vs. VGJ duty cycle

Fig. 13 Instantaneous jet exit blowing ratios a) and loss coefficient b) for various
duty cycles, all at 10 Hz. Velocity data a) taken with sub-miniature hotfilm probe in
VGJ exit at 63%Cx. Re = 25,000 (Reprint from Bons et al., 2001a).

multiplier of the beneficial jet influence. Because this relaxation time is essen-
tially constant for a given flow, reducing the pulse duty cycle can greatly increase
the free benefits of pulsed control. This behavior indicates that some economy of
jet flow is possible by optimizing the pulse duty cycle and frequency for a particu-
lar application.
To better understand the role of boundary-layer transition in the unsteady flow
control process, Reimann et al. (2006) acquired phase-averaged velocity and tur-
bulence data in the LPT separated shear layer with and without pulsing. The data
included a quantitative measure of intermittency to determine the state of the
boundary layer (i.e., laminar, transitional, or fully turbulent). Figure 14 shows the

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view

Fig. 14 Velocity data for no control (B = 0) on the Pack B at Re = 20,000. a) umean/Uin;


b) urms/Uin (%); c) skewness, and d) intermittency.
TURBOMACHINERY APPLICATIONS 275

no-control velocity data in a blade-fitting coordinate system (e.g., y is the direc-


tion normal to the wall at the x location along the blade from x/Cx = 0.3 to 1). The
urms/Uin data in Fig. 14b allow identification of the laminar boundary-layer separa-
tion location at 68% Cx and the free shear layer breakdown at 84% Cx. The skew-
ness data (Fig. 14c) show regions of reverse flow (high positive skewness) in the
separation bubble as well as a region of negative skewness along the transition
line. The transition line is also apparent from the intermittency plot (Fig. 14d).
When the VGJs are activated, the separation bubble experiences a cycle of
reattachment because of the disturbance, followed by a resurgence to steady-state
before the next jet pulse. This cyclic behavior can be viewed most efficiently in
a time–space plot, such as that shown in Fig. 15. The time–space plot shows data
acquired at a specific wall distance (y/Cx = 0.036) over 100 full cycles of jet actua-
tion. The 100 cycles are ensemble averaged to produce the two repeated cycles
shown in the figure. Since this wall distance is roughly 8 mm, the jet event (at
t = 0 and x/Cx = 0.6) does not become evident until t/T = 0.1. The duty cycle for
the VGJ was 25%, so the region of elevated turbulence and high intermittency
that extends upstream is a result of the jet disturbance. The jet convects down-
stream and interacts with the separation bubble at x/Cx  0.85. Once the VGJ
disturbance merges with the separation bubble, the speed of bubble motion slows
considerably (see second arrow in Fig. 15). A region of low urms fluid settles in
near t/T = 1 and x/Cx = 0.9. After some time, the separation zone expands
upstream (see third arrow) until it is again struck by the VGJ disturbance. The
intermittency plot (Fig. 15b) indicates that the boundary-layer transition location
cycles back and forth along the blade during the jet cycle. The dashed oval high-
lights a region of calm fluid that results after the passing of the highly turbulent
jet fluid. This calmed zone is similar to those identified by others following the
passage of an upstream wake disturbance (Gostelow and Thomas, 2003; Stieger
and Hodson, 2003).
LIVE GRAPH
Click here to view

Fig. 15 urms/Uin and intermittency time–space plots for Pack B. White vertical band
indicates transition location without control (B = 0) at this elevation (y = 8 mm,
see Fig. 14d).
276 H. F. FASEL ET AL.

To explore the role of the calmed zone further, an upstream wake generator
was added to the linear cascade by Reimann et al. (2007). The continuous reel
of cylindrical rods was located 12.7 cm (0.53Cx) upstream of the cascade inlet.
Unsteady wake disturbances were created using 6-mm diameter rods oriented in
the spanwise direction. The speed of the rods was adjusted to maintain a nor-
malized velocity of Urod/Uin = 1.05 (flow coefficient, Uin,axial/Urod = 0.85) with a
fluctuation of approximately ±2%. The period between rods was measured to be
225 ms, which is very close to the VGJ pulsing period of 200 ms used in Fig. 15.
The dimensionless forcing frequencies are both near F+ = 0.26. The rods were
spaced at L/S = 1.64, where L is the distance between the rods and S is the blade
spacing. The larger spacing between rods (compared to the cascade spacing)
was intended to simulate vane wakes impinging on a rotor blade row since the
vane count is typically 60–75% of the blade count for a given LPT stage.
Operating the wake generator without VGJ control, Reimann et al. (2007)
reported a similar (even larger) calmed zone in the region noted on the time–
space plot in Fig. 16.
The next obvious step was to coordinate the pulsed flow control with the unsteady
wakes in a constructive manner. To this end, Bloxham et al. (2007) synchronized
the VGJ pulse with the wake disturbance to minimize the separation bubble extent.
At optimal conditions, the jet disturbance arrived at the separation bubble just prior
to the breakdown of the wake-induced calmed zone. Consequently, the jet distur-
bance interacted with a smaller separation bubble. This resulted in the most sub-
stantial reduction of the separation region. Figure 17 shows the time–space plot for
this optimized condition. The bottom arrow denotes the calm region following the
wake disturbance while the top arrow marks the jet-induced calm region.
In order to quantify the effect of flow control with wakes, particle image veloci-
metry (PIV) data was used to identify the shape and size of the separation zone. The
LIVE GRAPH
Click here to view

Fig. 16 Time-space contour plot of urms/Uin for wake disturbances only at y = 6.2 mm
from blade surface.
TURBOMACHINERY APPLICATIONS 277

LIVE GRAPH
Click here to view

Fig. 17 Time–space contour plot of urms/Uin for pulsed VGJ synchronized with wake
disturbance. Data taken at y = 2 mm from blade surface.

total volume of low-momentum (separated) fluid was integrated and normalized by


the no control case. Figure 18 is a plot of this integrated measurement plotted as a
function of non-dimensional time (wakes with jets and wakes only data).
Figure 18 shows the impact of each of the disturbances and their relative effective-
ness in suppressing the separation bubble. The configuration with wakes only causes

1
Normalized Separation Zone Size

0.8

0.6

0.4

wakes/jets
0.2
wakes only

0
LIVE GRAPH 0 0.2 0.4 0.6 0.8 1
Click here to view t/T

Fig. 18 Integrated separation bubble size as a function of dimensionless time. The


data were normalized by the size of the no control separation bubble. (Reprint from
Bloxham et al., 2007).
278 H. F. FASEL ET AL.

a decrease in the normalized separation zone from 0.94 to 0.72. At t/T = 0.78 the
normalized separation bubble grows to nearly 0.81 as the 2D wake disturbance
impacts it. The bubble size then decreases to 0.58 as the 2D disturbance is ejected
from the blade. The average size of the separation bubble decreases very rapidly as
evidenced by the slope of the line during wake-induced control. A slower reduction
is noted in the VGJ-induced control. A comparison of the speed and size of these
reductions indicates that the spanwise-average wake-induced control might actually
have more impact than the jets. After the wake passes, the jet disturbance interacts
with a partial separation bubble. The remainder of the low momentum fluid is reen-
ergized, further decreasing the separation bubble to 0.42 (0.3 less than the wakes-only
configuration). These results suggest that at the optimal synchronizing configuration
the wake disturbance prepares the separation bubble for maximum jet effectiveness.
The same full-field PIV data that was used for the estimate of the separation
bubble size was also interrogated to determine whether jet-induced streamwise
vorticity was still evident in this more complex flowfield. Figure 19 contains
streamwise vorticity data (y–z plane) collected 5 ms before the VGJ deactivated. In
this study, the blowing ratio of the VGJ was B = 2, the jet duration was 50 ms, and
the duty cycle was 25%. The streamwise vorticity for four x/d locations (x/d = 10,
20, 25, and 35) is provided in the figure to track the vortical development. The VGJ
location is represented in the figure by the black arrow near z/d = 9 (jet hole center).
The plot of x/d = 10 depicts strong positive and negative VGJ-induced vorticity
cores. The cores are positioned near z/d = 7 and y/d = 2. These strong vorticity
cores dissipate in the subsequent plots. Despite the energy dissipation the positive
vortex maintains its structure up to x/d = 35 (well into the separation region). As
the vortical structures move downstream, they migrate away from the wall. By
x/d = 35, the positive core has migrated out to y/d = 4. Close examination of the full
PIV data set suggests that the vortex migrates away from the wall due to the pres-
ence of the separation bubble. The in-plane PIV data for subsequent time steps (not
presented) show that the vortex cores migrate back toward the wall as the separa-
tion bubble is reenergized and pushed off the turbine. The vortex cores also migrate
away from the jet location in the spanwise direction. This movement was expected
since the VGJ is injected with spanwise momentum. By x/d = 35, the positive
vortex core is positioned near z/d = 5. Given that vortical structures promote mix-
ing, it should be expected that the separation bubble would react to the presence of
the vortex. Close inspection of the 3-D nature of the VGJ’s impact on the upstream
end of the separation bubbles shows that reattachment begins near z/d = 5 and then
propagates outward. The downwash of the vortex causes the depression in the sepa-
ration bubble as high momentum fluid is carried into the low momentum bubble.
Similar VGJ-induced boundary-layer modifications have been observed by Hansen
and Bons (2006) and Khan and Johnston (2000). Although the in-plane PIV data
were collected without the addition of passing wakes, similar 3-D structures were
seen in both sets of data. These data suggest that streamwise vortices also partici-
pate in the removal of the separation bubble.

E. Active Flow Control as a Low-Pressure Turbine Design Variable


While significant effort has been focused on exploring and understanding the
fundamental physics of LPT flow control, ultimately the objective is to incorporate
TURBOMACHINERY APPLICATIONS 279

Fig. 19 Streamwise vorticity contours taken from three-component PIV data show-
ing jet-induced vorticity. VGJs located at x/d = 0 and z/d = 9 (hole center). Blowing
ratio, B = 2.

flow control into the turbine design process to enhance performance. As such,
several exploratory studies have been conducted to investigate design enhance-
ments that could be enabled through the use of flow control. The well documented
successes of VGJs at correcting off-design (low Re) separation deficiencies of an
existing LPT airfoil have led to the consideration of broader applications for VGJs.
For instance, VGJs could be used to design a more aggressive airfoil that would
otherwise be massively separated over its entire operating range. Specifically,
VGJ flow control could be used to obtain the same blade loading with reduced
axial chord, increase blade loading at constant chord and solidity, or decrease
solidity at constant blade loading. As such, the liberal use of VGJs in blade design
could potentially pave the way to savings in engine weight and part count without
a loss in stage efficiency.
280 H. F. FASEL ET AL.

The feasibility of this innovative use of VGJs was demonstrated by Sondergaard


et al. (2002b). In this study, the blade pitch (spacing) of an existing linear cas-
cade was increased at constant axial chord for a Reynolds number of 50,000.
(The Reynolds number is defined using the cascade inlet velocity and blade
axial chord. The equivalent Re based on design exit velocity is 82,000.) At this
Reynolds number, the unmodified cascade operated with a small separation
bubble on the suction surface. Increasing blade spacing effectively reduced the
cascade solidity and produced massive (non-reattaching) separation at the
design Reynolds number (with a six-fold increase in losses). This was attributed
to the larger peak pressure coefficient and the increased suction-side surface
distance associated with the unguided portion of the passage. Then, by strategi-
cally incorporating VGJs on the blade suction surface, the pitch-averaged blade
losses were reduced down to the standard pitch level. Figure 20 shows a plot of
wake pressure loss coefficient g vs VGJ blowing ratio B at a Reynolds number
of 50,000. At the design pitch S = S0 the flow was attached at the Reynolds
number of 50,000, and the loss coefficient for B = 0 was just over 0.1. The effect
of blowing was insignificant until the blowing ratio exceeded 2, at which point
the VGJs caused the loss to increase. This was due to the fact that once the VGJs
became strong enough they blew the attached boundary layer off and created
additional losses.
Increasing the pitch to 125% of the design pitch (with B = 0) caused a signifi-
cant increase in loss coefficient to approximately five times the corresponding
value for the design pitch level. This was due to the increase in uncovered turn-
ing, which resulted in a large separation on the suction side of the blades.
LIVE GRAPH
Click here to view
0.7
Re 50k, Tu 1% Ref
0.6 S = So
S = 1.25 So
Pitch-Normed Loss Coeff

S = 1.5 So
0.5
S = 1.5 So, 2 row
S = 2 So
0.4

0.3

0.2

0.1

0
0 1 2 3 4 5
B

Fig. 20 Wake pressure loss coefficient vs blowing ratio, B. Re = 50,000 and


FSTI = 1%. VGJ injection at 45% Cx (45% and 63% injection for 2-row). Four pitch
spacings. Dotted line represents reference loss level, S = So and B = 0 (Reprint from
Sondergaard et al., 2002b).
TURBOMACHINERY APPLICATIONS 281

Application of VGJ blowing at 45% Cx significantly affected the loss coefficient.


For blowing ratios of approximately 1 or greater, the loss coefficient for the
125% pitch case was driven down to the design level. As the pitch increased, the
zero blowing loss level increased as the size of the separation zone increased. At
150% of the design pitch, the zero blowing loss had increased to approximately
six times the loss level for the design pitch. Again, it was possible to reduce the
loss to design levels with the application of VGJ blowing. The loss was reduced
to only twice the design level at a blowing ratio of B = 2, and down to the design
level by a blowing ratio of between 4 and 4.5. Increasing the pitch to 200% of
the design pitch further increased the size of the separation zone and increased
the zero blowing loss to nearly seven times the design value. Again the loss was
driven down significantly with the application of VGJs, though in this particular
instance the VGJ feeder system used was not able to provide sufficient mass
flow to determine if the losses could be driven down to the design level. The
implication of Fig. 20 was that equivalent cascade performance could be attained
with up to 50% fewer blades.
While this experiment established the feasibility of the integrated VGJ design
approach, it would be premature to assume that the full potential of the new tech-
nology has thus been realized. In a related experiment by Merchant et al. (2004),
researchers demonstrated the beneficial effects of applying boundary-layer suction
to compressor blading to reduce wake losses. One important finding from their
work was that retrofitting an existing compressor airfoil with boundary-layer suc-
tion did not necessarily improve the stage performance. Indeed, the full potential
of the “aspirated compressor” was only realized when an entirely new airfoil was
designed incorporating the suction slot. Likewise, it is expected that the full poten-
tial of VGJs in LPTs will only be achieved by integrating the flow control devices
into the initial airfoil design process. In a follow-on study by Bons et al. (2005), a
re-design of the Pack B airfoil was performed using the airfoil design, analysis, and
optimization system implemented recently at AFRL. Recently published transition
models (Praisner and Clark, 2004) and high-lift airfoil design information (Praisner
et al., 2004) were used in conjunction with MISES and the flow solver of Dorney
and Davis (1992) to define an airfoil with a balanced loading distribution and 17%
more lift than the original shape. The airfoil was designed to incorporate VGJs just
upstream of the predicted separation location at low Reynolds numbers. The result
of this design process is the so-called L1M blade profile. Unexpectedly, the new
L1M blade profile was found to be resistant to un-reattached boundary-layer sepa-
ration at inlet Reynolds numbers below 15,000. Thus, flow control was not criti-
cally needed to reach performance objectives. At the same time, significant
separation bubbles were predicted to occur at inlet Reynolds numbers of 20,000
and 50,000, and it was shown experimentally that these separations were effectively
controlled with pulsed-blowing from VGJs at these conditions. Efforts are under-
way to push the blade performance to the brink of current design codes and beyond
using flow control to stave off separation losses as far as possible.

F. Low-Pressure Turbine Separation Control using Plasma Actuators


Dielectric barrier discharges (DBD) have been used to effectively reattach low
Reynolds number flows in LPT Pack B cascades and simulated Pack B blading at
282 H. F. FASEL ET AL.

Notre Dame (Huang et al., 2002), NASA Glenn (Hultgren and Ashpis, 2003), the
Air Force Academy (List et al., 2003), and the Air Force Research Laboratory (Boxx
et al., 2006a, b). Aerodynamic applications include the work of the previous research
groups as well as the University of Tennessee and NASA Langley (Roth et al.,
1998), the University of Kentucky/Texas A & M (Jacob et al., 2004), and many
others. DBDs consist of a pair of electrodes separated by a dielectric material. When
sufficient alternating current (ac) voltage is applied, a plasma is formed between the
two electrodes and a volume force is exerted on all charged particles (ions and
electrons) in the direction of the gradient of the square of the electric field. Both
signs of charge receive a body force in the same direction. Most of the experiments
in the literature for aerodynamic applications involve high-frequency 3000–
10,000 Hz ac discharges at 3000–10,000 V. Examples of high-frequency and pulsed
direct current (dc) discharges from AFRL experiments for separation control of
“simulated” separated Pack B flows will be presented. For these experiments the
typical discharge volume was ~4 mm long × 4 mm high × electrode width.
In the experiments with DBD flow control, the Pack B blade suction surface
pressure distribution was replicated on a flat plate which was located in the center
of a rectangular test section. This was accomplished by modifying the shape of the
upper wall of the test section. The DBD device was located at the (simulated) 55%
chord location and aligned parallel to the span. Flow separation from the upper
channel wall was prevented by applying suction through a 3-cm wide slot. This
setup allowed for simulations of adverse pressure gradients that were similar to
the Pack B pressure gradient as well as much more aggressive pressure gradients.
In Fig. 21 the simulated high Reynolds number wall pressure distribution is com-
pared with a low Reynolds number distribution (with and without (closed valve)
suction applied at the upper wall) for Re = 25,000. The low Reynolds number
distribution (open symbols) shows flow separation which was found to be more
severe and earlier than in full cascade experiments.
Using PIV, velocity, turbulence, and vorticity data were obtained for a high-
frequency ac discharge with 0, 15, 20, and 25 watts of electric input power.
Without actuation (Fig. 22a) the flow separates approximately 10 mm upstream of
the DBD electrodes. At about 10 mm downstream of the actuator, the wall-normal
extent of the separated flow region is about 5 mm. With actuation, for an electrical
input power of 15 W, the extent of the separated flow region is already reduced
(Fig. 22b). At 20 W input power the separated flow region is virtually eliminated
and the velocity profiles show first indications of a developing wall jet (Fig. 22c).
At 25 W flow separation is eliminated upstream of the electrodes and a wall jet is
clearly defined (Fig. 22d). A small separated flow region remains upstream of the
electrodes which in this case can be related to a weak remaining gradient of the
electric field in upstream direction. Wall jet velocities are typically in the order
of ~1 m/s.
Profiles of turbulence intensity (RMS of velocity fluctuations) obtained from
the PIV measurements are shown in Fig. 23. These profiles were obtained by
averaging over 1000 measurements which were taken at a sampling rate
of ~6000 Hz for each of the measurement locations. The actuator influences the
turbulence intensity profiles both upstream and downstream of the actuator loca-
tion. As the actuator amplitude is increased, the peak of the turbulence intensity is
drawn closer to the wall and a large secondary peak is induced near the wall.
LIVE GRAPH TURBOMACHINERY APPLICATIONS 283
Click here to view

Fig. 21 Wall pressure distribution. Flat plate experiments with imposed pressure
distribution of Pack B blade suction surface. Comparison of high and low Reynolds
number data.

For the same flat plate experiment with Pack B pressure distribution the actua-
tor was also powered by a pulsed dc discharge. Pulsed dc discharges with a pulse
length of 22 ¥ 10–9 s (Rivir et al., 2004) to 2 ¥ 10–6 s have been investigated at
AFRL (Fig. 24). Compared with the ac operation, with pulsed discharges, higher
voltages and stronger electric fields can be obtained, resulting in comparatively
more power per pulse as well as lower average power consumption for low duty
cycles. With pulsed dc discharges a peak power of 7–60 kW per pulse was mea-
sured. The average power consumption was 0.2–0.05 W compared to the average
5–25 W power consumption in case of the high-frequency ac operation. Pulsed dc
actuation also offers the possibility of coupling into higher modes in the flow—
and this is particularly important as DBDs are applied to higher velocity flows.
The question remains whether, for the small amount of average power fed into
the actuator for pulsed dc discharge operation, the actuation has an effect on the
separated Pack B flow. Shown in Fig. 25 are PIV measurements without flow
control and with pulsed dc discharge actuation with ~7 kW peak per pulse
(Fig. 24), a pulse length of 10–6 s, and 100 pulses/s. The flow visualizations clearly
show that the flow was reattached over the actuator when the actuator was on.
Also, with actuation the velocity profiles indicate a 60% reduction in the separation
18 mm downstream of the actuator. Very short pulsed dc discharges still have a
large effect on the separated flow region.
In summary, DBDs have the capability to reattach separated low Reynolds
number flows at pressure gradients and operating conditions that are typical of
284 H. F. FASEL ET AL.

Fig. 22 Velocity isocontours and velocity vectors for separating boundary layer with
Pack B pressure distribution and DBD actuator located near x = 0. Actuator off and
actuator on with actuator power settings of 15, 20, and 25 W (Reprint from Boxx et al.,
2006a, b).

low Reynolds number LPT flows. The induced changes in the wall velocities that
can be obtained with DBDs are limited to typically 1–10 m/s. The range of pos-
sible pulse lengths and duty cycles makes DBDs applicable for a wide range
of flows. In particular, the wide range of possible DBD operating parameters
allows the finding of a setting where the flow responds to the actuation (such that
LIVE GRAPH TURBOMACHINERY APPLICATIONS 285
Click here to view

X = –6.3 mm X = –4.1 mm X = –1.9 mm X = –0.1 mm


15

10

5
Y- location (mm)

0
X = 2.6 mm X = 4.8 mm X = 7.1 mm X = 9.3 mm
15

10

0
0 0.2 0.4 0 0.2 0.4 0 0.2 0.4 0 0.2 0.4
Turbulence internsity

Fig. 23 Profiles of turbulence intensity obtained from PIV for Re = 16,400 at stations
upstream and downstream of actuator. Circles: actuator off; crosses: actuator on,
15 W; triangles: 20 W; diamonds: 25 W (Reprint from Boxx et al., 2006a, b).

LIVE GRAPH
Click here to view

Fig. 24 Current, voltage, and power for single DC discharge pulse.


286 H. F. FASEL ET AL.

Fig. 25 Streamlines without and with pulsed DC discharge DBD actuation (Reprint
from Wall et al., 2007).

a coupling of the flow dynamics and the actuation is achieved). It also allows for
the possibility of accomplishing this coupling in a beneficial nonlinear fashion.

III. Simulations
A. Overview
Before computational fluid dynamics (CFD) can be employed for investigating
AFC for LPT applications the applicability of the chosen specific numerical
method for LPT simulations needs to be demonstrated. Particular challenges for
CFD are the transitional nature of the flow and associated resolution requirements.
The relevant Reynolds number range is barely accessible for direct numerical
simulations (DNS) while the transitional nature of the flow precludes the use of
common turbulence modeling approaches. Reynolds averaged Navier–Stokes
(RANS) coupled with an intermittency function (e.g., Suzen and Huang, 2005)
show some potential. Large-eddy simulation (LES) (e.g., Michelassi et al., 2003)
and implicit LES (ILES) (e.g., Raverdy et al., 2003; Rizzetta and Visbal, 2003a,
2004, 2005; Gross and Fasel, 2005a, b), where the diffusion of the numerical
scheme provides the model contribution of typical LES sub-grid stress models
(Margolin and Rider, 2002), are considerably more accurate. High fidelity results
can only be obtained from true DNS which do not rely on any sort of turbulence
modeling. For example, using DNS, Wu and Durbin (2001) showed that turbulent
wakes can result in by-pass transition and thus suppress separation on the suction
side of the T106 LPT blade. Kalitzin et al. (2003) and Wissink and Rodi (2006)
showed that large FSTI can also accelerate transition. Both DNS and LES appear
to be appropriate tools for studying LPT flows.
Three-dimensional ILES of a Pack B blade, carried out by Rizzetta and Visbal
using a high-order accurate compressible code, focused on understanding the fluid
dynamics of separation control by pulsed VGJs (Rizzetta and Visbal, 2003a, 2004,
2005) and plasma actuators (Rizzetta and Visbal, 2007). For these investigations
TURBOMACHINERY APPLICATIONS 287

Rizzetta and Visbal placed an embedded grid over the VGJ holes and the plasma
actuators for increasing the local grid resolution. Pulsed VGJs were found to result
in an earlier transitioning of the suction side boundary layer. A phenomenological
model was employed to represent the plasma induced volume force on the flow.
Pulsed plasma actuation was found to be more effective than continuous opera-
tion. For pulsed actuation, counter-flow actuation was found to be more effective
than co-flow actuation. The effectiveness of the latter was attributed to the genera-
tion of spanwise vortical structures that entrain high-momentum fluid from the
free-stream. The same geometry was also investigated numerically by Gross and
Fasel (2005a, b) and Postl et al. (2003, 2004). High-resolution DNS of laminar
separation bubbles on a flat plate at LPT conditions, a related model problem
which is very helpful at elucidating the physics of separation control by VGJs,
were performed by Postl (Postl, 2005; Postl et al., 2003, 2004; Wissink and Rodi,
2006). Wissink and Rodi studied the effect of an oscillating uniform inflow or a
uniform inflow with or without turbulent free-stream fluctuations. In the first case,
the separated boundary layer was found to “roll up” due to a Kelvin–Helmholtz
instability followed by a rapid transition to turbulence. With turbulent fluctuations
the Kelvin–Helmholtz instability was triggered much earlier, resulting in a drastic
reduction of the size of the separation bubble.
Our efforts focused on extracting the physical mechanisms responsible for the
astonishing effectiveness of steady and pulsed VGJs. This will be discussed in
more detail in later sections. We have taken a two-pronged approach towards
uncovering the fundamental mechanisms for AFC for LPT blades. In the first
approach (Sec. B), simulations were performed for the flow through the entire
LPT cascade, investigated experimentally by Rivir and co-workers at AFRL (Bons
et al., 2001a, 2002; Sondergaard et al., 2002a) where we investigated separation
control by pulsed VGJs and pulsed blowing through a slot. In the second approach
(Sec. C), we used a model geometry, a laminar boundary layer on a plate (with and
without curvature), and imposed the same streamwise pressure gradient as mea-
sured in the experiments for a turbine blade. This simpler geometry allowed us to
focus available computational resources on the separated flow region and to
resolve all relevant scales of motion, from the laminar to the turbulent flow regime,
which enabled us to extract the relevant physics associated with separation control
by steady and pulsed VGJs.

B. Approach 1: Numerical Simulations of the Pack B and L1M


Low-Pressure Turbine Blades
1. Simulation Details
For computations of entire LPT blades we employed a finite volume code based
on the compressible Navier–Stokes equations in curvilinear coordinates. The
convective terms were approximated with fifth-order-accurate and ninth-order-
accurate upwind schemes based on a weighted essentially non-oscillatory extra-
polation of the characteristic variables and the Roe scheme (Gross and Fasel,
2002). A fourth-order-accurate discretization was employed for the viscous terms.
The governing equations were advanced in time with a second-order-accurate
implicit Adams–Moulton method. For ILES, the convective terms were discretized
with a second-order-accurate total variation diminishing scheme and the viscous
288 H. F. FASEL ET AL.

terms were computed with a second-order-accurate discretization. For the flow


simulation methodology simulations (FSM) (Fasel et al., 2002; Speziale, 1998),
the 1998 k-w turbulence model equations (Wilcox, 2000) were solved using a
second-order-accurate discretization and the Reynolds-stresses were computed
from the explicit algebraic stress model by Rumsey and Gatski (2001).
The LPT geometries chosen for the current studies are the Pratt and Whitney
Pack B LPT blade and the L1M blade (Bons et al., 2005). Both blades have an
inflow angle of 55 deg and a design exit angle of 30 deg (measured from the plane
of the cascade). Our Pack B simulations were set up according to the experiments
by Bons, Sondergaard, Rivir and co-workers (Bons et al., 2001a, 2002, 2005;
Sondergaard et al., 2002a) where upstream wakes, surface roughness, rotation,
and 3-D effects were not considered. The experimental cascade consisted of eight
blades with a span of 5Cx. The Pack B simulations were carried out for a
Reynolds number based on axial chord and inlet velocity vin of 25,000. The ratio
of blade spacing S and axial chord Cx was 0.88 for the original cascade of Rivir
and co-workers and 1.1 for a cascade with 25% larger blade spacing. The L1M
simulations were set up according to experiments by Bons et al. (2005). The
experimental L1M cascade consisted of three blades. The simulations were for a
Reynolds number of 20,000. The blade spacing was S/Cx = 1.01 (as in experi-
ments). The FSTI in the L1M experiments was about 3% (Bons et al., 2005). In
our LPT simulations, the FSTI was 0% (laminar inflow).
The computational grids for our simulations of the entire LPT cascade consisted
of five blocks (Fig. 26). Grid points were concentrated in the separated flow region
on the suction side of the blade and in the wake. The no-slip condition was applied
at the wall except at the actuator location where a forcing velocity was prescribed.
The wall was treated as adiabatic. A characteristics-based boundary condition was
applied at the inflow and outflow boundaries (Gross and Fasel, 2007b). Periodicity
conditions were employed at all other boundaries. In computations by Rizzetta
and Visbal (2003a, b, 2004, 2005) a spanwise extent of the computational domain
of 0.2Cx was found to be sufficient for capturing the relevant fluid dynamics. We

Fig. 26 Computational grid for Pack B LPT simulations.


TURBOMACHINERY APPLICATIONS 289

Fig. 27 Schematic of forcing through slot.

chose the same spanwise grid extent for our 3-D simulations. Further details can
be found in (Gross and Fasel, 2005a, b, 2007a).
Bons, Sondergaard, Rivir and co-workers employed VGJs for controlling separa-
tion from the LPT blade (Fig. 4). Compressed air is issued through holes that are
inclined (pitch and skew angle) with respect to the flow direction. The VGJs were
located slightly upstream of the separation line and distributed over the entire span
with constant spanwise spacing. Pulsed actuation (Fig. 12), the so-called reduced
duty cycle forcing (Bons et al., 2001a, 2002), was found to be more effective than
steady VGJs (Sondergaard et al., 2002a). The pulsed in-phase actuation of the VGJs
introduces a 2-D disturbance component. The discrete spanwise spacing of the VGJs
additionally introduces a 3-D disturbance component with a spanwise wavelength
that is identical to the spanwise spacing of the VGJs. In the simulations, the VGJs
are modeled by prescribing the jet exit velocities on the surface of the blade.
In both, experiments (Huang et al., 2006b) and simulations (Rizzetta and Visbal,
2007) the flow was also controlled using plasma actuators. Plasma actuators intro-
duce a weak spanwise constant disturbance input, the effectiveness of which
depends on the configuration of the actuator. A spanwise constant disturbance
input can also be accomplished by blowing (and/or suction) through a spanwise
slot (Fig. 27). In our simulations the slot was modeled by prescribing a time-
dependent wall normal velocity on the surface of the blade.

2. Uncontrolled Flow
a. Pack B Geometry. Prior to our AFC simulations we first validated our
code for the Pack B blade without flow control (Gross and Fasel, 2005b).
Instantaneous contours of spanwise vorticity

∂v ∂u
wz = - (5)
∂x ∂y

for the design blade spacing S/Cx = 0.88 are shown in Fig. 28. The laminar boundary
layer separates approximately at the beginning of the uncovered turning. As a result
of hydrodynamic instabilities of the separated boundary layer and the bubble
(inflectional velocity profile, Kelvin–Helmholtz instability, possibly global instabil-
ity) disturbances are strongly amplified, leading to a “roll-up” of the separated
boundary layer and the formation of spanwise vortical structures. These structures
290 H. F. FASEL ET AL.

Fig. 28 Iso-contours of spanwise vorticity (uncontrolled flow, Pack B, S/Cx = 0.88).


a) 2-D simulation and b) 3-D simulation. (Reprint from Gross and Fasel, 2005b)

increase wall normal mixing and cause flow reattachment in the mean. All 3-D
motion is suppressed in the 2-D simulation and the intensity of the 2-D structures is
over-predicted when compared with the 3-D case. In the 3-D simulations, the flow
is allowed to transition. The eddy-viscosity introduced by small-scale turbulent
structures weakens the coherence of the 2-D structures thereby delaying reattach-
ment. Since the disturbances that seed the spanwise structures were not introduced
deliberately into the flow and since the wake shedding was self-sustained it is pos-
sible that the flow is absolutely unstable. Theofilis and co-workers applied a BiGlobal
instability analysis tool to identify a global instability of the separation bubble on
the suction side of an LPT blade at relatively low Reynolds numbers (Abdessemed
et al., 2004; Theofilis and Sherwin, 2004) in addition to the Kelvin–Helmholtz and
Tollmien–Schlichting mechanisms. By analysis of a 2-D time-dependent base flow
obtained from a DNS using Floquet theory, 3-D instability modes were identified
(Abdessemed et al., 2006). These modes were found to affect both the bubble and
the wake simultaneously. Clearly, global modes have to be considered when optimal
frequencies and spanwise wavelengths for AFC schemes are sought.
Time-averaged results for the wall pressure coefficient cp, are shown in Fig. 29.
Also included in this figure are measurements by Bons et al. (2001a), Sondergaard
et al. (2002a), and Huang et al. (2006a) as well as results from a 3-D ILES by
Rizzetta and Visbal (2003b). The FSTI was 0.08% in the experiments by Huang
et al. and about 1% for the other experiments. Compared to the experimental data,
the 2-D results show a somewhat later separation and a slightly smaller pressure
plateau in the separated region. The hump near the TE for the 2-D calculation
which is not visible in the 3-D simulations appears to be an artifact of the overly
strong 2-D structures.
Time-averaged wall normal profiles of total velocity averaged in the spanwise
direction and normalized by the local velocity maxima are shown in Fig. 30.
The computed results are compared with experimental data (Bons et al., 2001a;
Sondergaard et al., 2002a). At 68% axial chord, the computed velocity profiles are
slightly less full than the measured profiles indicating that the flow is closer to
separation. Downstream, at 77%, the computed velocity profiles obtained from
TURBOMACHINERY APPLICATIONS 291

LIVE GRAPH
Click here to view

Fig. 29 Wall pressure coefficient for Pack B geometry at S/Cx = 0.88. (Reprint from
Gross and Fasel, 2005c)

the 3-D simulations match the experimental data quite well. The bubble thickness
is under-predicted in the 2-D simulations, indicating mixing that is too strong
because of the high intensity 2-D structures. The thickness of the separation
bubble is underpredicted at 84 and 92% axial chord. A fundamental difference in
the setup between the simulations and the experiments may be the reason for
the observed differences. This may include effects of FSTI (neglected in the
simulations), or differences in the inflow and outflow angles, effects of minor
geometric variations of the blade geometries, etc.
A visualization of the instantaneous data obtained from the 3-D DNS is shown
in Fig. 31. For this figure the integration domain was repeated once in the spanwise

LIVE GRAPH
Click here to view

Fig. 30 Wall-normal velocity profiles (Pack B, S/Cx = 0.88).


292 H. F. FASEL ET AL.

Fig. 31 Iso-surface of Q = 1 (Pack B, S/Cx = 0.88). Perspective view near TE.

direction so that the developing flow structures can be observed better. Shown
in Fig. 31 are iso-surfaces of the Q-criterion (Hunt et al., 1988). A positive
Q-criterion indicates areas where rotation dominates strain. The 2-D spanwise
structures that periodically develop as a consequence of a shear layer instability
become unstable with respect to 3-D disturbances. These disturbances initially
cause a 3-D modulation of the 2-D structures, and eventually lead to a rapid
breakdown to turbulence upstream of the TE. Although the wake is turbulent,
the unsteady “footprints” of the 2-D shear layer instability can still be recog-
nized by the presence of distinct spanwise coherent structures in the wake. With
a grid resolution of 2.8 million cells the current DNS results are still under-
resolved. Rough estimates indicate that the grid resolution has to be increased
by at least one order of magnitude for resolving all turbulent length scales down
to the Kolmogorov length scale. Strictly speaking, the current DNS is therefore
still an ILES (Margolin and Rider, 2002). However, in this chapter the abbrevia-
tion ILES is employed solely for simulations carried out with the second-order-
accurate discretization.
b. L1M Geometry. Results obtained from a 2-D FSM simulation of the
uncontrolled L1M flow are shown in Fig. 32 (Gross and Fasel, 2007a). The
downstream extent of the separated flow region becomes visible when consider-
ing isocontourlines of the streamfunction of the time-averaged flow data
(Fig. 32b). The flow reattaches (in the mean) well upstream of the TE of the
blade. The picture changes considerably when 3-D motion is allowed (Fig. 33).
The flow separates earlier at 55% axial chord (57% for 2-D case). A secondary
instability mechanism leads to the amplification of 3-D disturbances resulting in
the appearance of 3-D structures near the TE and in the wake. The 3-D structures
weaken the coherence of the spanwise structures, resulting in an enlargement of
the separation bubble.
A comparison of the computed wall pressure coefficient with experimental data
and a design code (MISES) prediction (Bons et al., 2005) is shown in Fig. 34. The
TURBOMACHINERY APPLICATIONS 293

Fig. 32 2-D simulation (L1M): a) spanwise vorticity and b) streamfunction (Reprint


from Gross and Fasel, 2007a).

agreement between the 3-D simulation and the MISES prediction is adequate.
Again, a pronounced hump of the 2-D result near the TE can be detected which
can be attributed to overly strong spanwise structures. The later separation in the
2-D case causes the pressure in the separated flow region (pressure plateau) to be
lower than in the 3-D case. The significantly smaller size of the separation bubble in
the experiments (see Reimann et al., 2006, Fig. 5) can likely be attributed to the
FSTI of about 3% that was present in the experiments.

3. Open-loop Control (2-D)


a. Pack B Geometry. For investigating open-loop control of the Pack B
blade we increased the blade spacing by 25% (S/Cx = 1.1) to obtain a larger

Fig. 33 3-D simulation (L1M): a) spanwise vorticity and b) streamfunction (Reprint


from Gross and Fasel, 2007a).
294 H. F. FASEL ET AL.

LIVE GRAPH
Click here to view

Fig. 34 Wall pressure coefficient for L1M geometry at S/Cx = 1.01 (Reprint from
Gross and Fasel, 2007a).

separation bubble (Gross and Fasel, 2005b). Instantaneous spanwise vorticity


contours for the uncontrolled flow are shown in Fig. 35. With the larger blade
spacing, the separated region and hence the performance losses are much larger.
The area enclosed by the cp-curves (Fig. 36) is larger than for the design blade
spacing (Fig. 29) since each individual blade has to exert a larger aerodynamic
force on the flow. The hump in the cp-curve of the 2-D calculation is far less
pronounced for the 3-D simulation indicating significantly weaker spanwise coher-
ent structures when compared with the calculation for the design blade spacing.
Three-dimensional simulations are too expensive with currently available com-
puting resources for AFC parameter studies. Therefore, we conducted 2-D simu-
lations for investigating the response of the flow to a reduced duty cycle actuation.
The two-dimensionality of the simulations precludes any 3-D motion and therefore
alters the fluid dynamics to the extent observed in the previous comparisons
between 2-D and 3-D simulations. The results from this study are, nevertheless,
very helpful for explaining the flow response to a time-periodic actuation.

Fig. 35 Iso-contours of spanwise vorticity (Pack B, S/Cx = 1.1): a) 2-D and b) 3-D
results (Reprint from Gross and Fasel, 2005b).
TURBOMACHINERY APPLICATIONS 295

LIVE GRAPH
Click here to view

Fig. 36 Wall pressure coefficient for Pack B geometry at S/Cx = 1.1 (Reprint from
Gross and Fasel, 2005b).

Furthermore, in light of some of the 3-D results with 2-D disturbance input shown
later in this chapter, we concluded that investigating the response of the flow to a
2-D disturbance does not always require full-fledged 3-D simulations.
For the simulations discussed here, a forcing slot of width b = 0.01Cx was located
at 57% chord. This location was chosen based on the experimental observations.
Placing the actuator further upstream in the favorable pressure gradient region
would reduce its efficiency as (according to linear stability theory) disturbances
would experience a weaker amplification or might even be dampened. The slot was
resolved with three cells and a top hat velocity profile was applied over the width of
the slot. At this resolution, the near-slot fluid dynamics were not resolved. Instead,
the forcing merely introduced a 2-D disturbance into the flow. For the cases shown
here, a blowing ratio of B = 1 and a duty cycle of t = 10% were chosen. The momen-
tum coefficient, cm = (b/Cx)tB2, was 10-3. Instantaneous visualizations of the con-
trolled flow for two different forcing frequencies are shown in Fig. 37. Comparison
with Fig. 35 (left) reveals that the dynamics of the flow are changed profoundly.

Fig. 37 Iso-contours of spanwise vorticity (Pack B, S/Cx = 1.1): a) control with


F + = 2.5 and b) 6.67 (Reprint from Gross and Fasel, 2005b).
296 H. F. FASEL ET AL.
LIVE GRAPH LIVE GRAPH LIVE GRAPH
a) Click here to view b) Click here to view c) Click here to view
6 3 3
5
4 2 2
3
t

t
2 1 1
1
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x x

Fig. 38 Space/time diagrams of wall vorticity on suction side of blade (wz = -1200 . . .
1200) for a) uncontrolled case, b) controlled cases with F + = 2.5, and c) 6.67 (Pack B,
S/Cx = 1.1) (Reprint from Gross and Fasel, 2005b).

Figure 38 shows x/t diagrams of the wall vorticity for the uncontrolled case and
two controlled cases. Without flow control, vortices are shed in an irregular fash-
ion and the frequency spectrum (Fig. 39) is very broad. More energy is contained
in the lower-frequency range of the spectrum than for the case with design blade
spacing (Gross and Fasel, 2005b) since the separation bubble and the associated
length scales are larger. For the controlled cases, separation is reduced consider-
ably (Figs. 36 and 38). The shortest separated flow region was obtained when the
flow was forced with F+ = 6.67. For this forcing frequency the introduced distur-
bances are most amplified (Fig. 40). This frequency is above the frequency range
of the energy-containing structures of the uncontrolled flow, which is approxi-
mately 1–3 (Fig. 39).
The time-averaged ratio of the nondimensionalized normal and axial aero-
dynamic forces cy /cx is increased by as much as 19.4% for forcing with F+ = 6.67
(Fig. 41). A simple control volume analysis shows that this parameter is a measure
for the total flow turning (difference between cascade inflow and outflow angle)
and hence a performance index. When considering the individual aerodynamic
forces it becomes clear that the gain in cy/cx is mostly due to a reduction of cx. The

LIVE GRAPH LIVE GRAPH LIVE GRAPH


Click here to view Click here to view c) Click here to view
a) b)
20 20 20

15 15 15
F+

F+

F+

10 10 10

5 5 5

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x x

Fig. 39 Fourier transform of wall vorticity on suction side of blade (A(wz) = 0 . . . 400)
for a) uncontrolled case, controlled cases with b) F + = 2.5, and c) 6.67 (Pack B,
S/Cx = 1.1) (Reprint from Gross and Fasel, 2005b).
TURBOMACHINERY APPLICATIONS 297

LIVE GRAPH
Click here to view

Fig. 40 Amplitude of fundamental disturbance (forcing frequency) for three differ-


ent forcing frequencies (Pack B, S/Cx = 1.1). Amplitudes were obtained from Fourier
transform of wall vorticity (Reprint from Gross and Fasel, 2005b).

separation bubble is located near the TE of the blade in the area of uncovered
turning where the pressure recovery takes place. In the uncontrolled case, the flow
cannot negotiate the adverse pressure gradient and separates, resulting in a pres-
sure distribution below the design pressure in the aft part of the blade (Figs. 29
and 36). Because of the curved geometry of the blade, this loss shows up mainly
in the cx-coefficient. When the control is turned on, separation is reduced and the
pressure at the aft part of the blade is increased. When the flow “locks in” to the
forcing, as for example for F+ = 3.33 and F+ = 6.67, the dynamic aerodynamic
loads on the blade are increased considerably (Fig. 41). This can be avoided by
forcing with F+ = 5 (a frequency that is slightly lower than the most amplified
frequency). For this frequency the fundamental (F+ = 5) and higher harmonic
(F+ = 10) are both close to F+ = 6.67 and both are amplified. The loss in aero-
dynamic performance due to this compromise is small. The increase in cy /cx of the

Fig. 41 Time averages and standard deviations s of aerodynamic coefficients cx


and cy. 1/F+ = 0 result is for uncontrolled case (Pack B, S/Cx = 1.1) (Reprint from Gross
and Fasel, 2005b).
298 H. F. FASEL ET AL.

Fig. 42 Streamfunction isocontours for a) uncontrolled case, controlled cases with


b) F + = 2.5, and c) 6.67 (Pack B, S/Cx = 1.1). (Reprint from Gross and Fasel, 2005b).

controlled cases is associated with a larger circulation and a smaller time-averaged


separation bubble (Fig. 42) when compared with the uncontrolled case.
b. L1M Geometry. Using 2-D simulations we explored blowing ratios in
the range B = 0.01–1 and forcing frequencies in the range F+ = 1–10 in a system-
atic manner and tracked the performance index cy /cx (Fig. 43) for harmonic
blowing through a slot of width 0.01Cx which was located at 50% axial chord
(Brehm et al., 2006). In Fig. 43 a clear maximum of the performance index can
be found for B  0.1 and F+  7. The gain in cy /cx from the uncontrolled
(cy /cx = 1.71) to the controlled case (cy /cx = 2.12) is 24%. The frequency range
for which the flow control is beneficial becomes largest for B = 0.1–0.2. As the
blowing ratio is increased the control slowly loses effectiveness. For larger
blowing ratios the boundary layer separates from the wall at the blowing slot
location. This is counterproductive: although the control still leads to an earlier
reattachment (when compared to the uncontrolled case) the flow separates
earlier. For high-enough blowing ratios, the jet acts as a flow obstruction.

Fig. 43 Dependence of performance index cy /cx on blowing ratio B and forcing


frequency F+ (L1M) (Reprint from Brehm et al., 2006).
TURBOMACHINERY APPLICATIONS 299

4. Open-Loop Control (3-D)


The parameter study indicated that the flow amplifies 2-D disturbances and
that the optimal forcing frequency and blowing ratio were F+  7 and B  0.1.
We conjectured that if this flow control scheme worked for the 2-D case it would
also be somewhat effective for the 3-D case (Gross and Fasel, 2007a). Following
numerous LPT experiments (e.g., Bons et al., 2005; Reimann et al., 2006) it was
decided to investigate open-loop flow control using pulsed VGJs (PVGJs) as in
the experiments. The jet exit holes were positioned at x = 0.56Cx and had a
30 deg pitch and 90 deg skew angle (see Fig. 4). The hole spacing was identical
to the spanwise extent of the computational domain, Dz = 0.2Cx. Assuming fully
developed laminar pipe flow in the feedlines a parabolic velocity distribution
was prescribed over the holes. Each hole was resolved with 7 × 7 cells and had
a diameter of d = 0.0388Cx. As in the 2-D simulations discussed earlier for
open-loop control, a forcing frequency of F+ = 7 was chosen. The duty cycle
was t = 10% and the blowing ratio was B = 4, resulting in a momentum coeffi-
cient of cm = 4.73 × 10-3. For comparison, open-loop control using harmonic
blowing through a slot with F+ = 7 and B = 0.1 was also investigated. The slot
was located at x = 0.55Cx and had a width of b = 0.015Cx. The momentum
coefficient was cm = 5.63 × 10-5.
A comparison of both controlled cases (Fig. 44) with the uncontrolled flow
(Fig. 33) shows that both flow control strategies effectively reduce the size of the
separated flow region. The various smaller size separation bubbles seen for the
case with PVGJ control may be an artifact of the short time-averaging interval.
When considering the spanwise vorticity it appears that flow control by PVGJs
results in an earlier transitioning of the flow. The attached turbulent boundary
layer can withstand a larger adverse pressure gradient. On the other hand, control
by harmonic forcing through a slot appears to weaken the amplification of 3-D
structures. In this case, the flow can be seen to regularize, it becomes almost per-
fectly 2-D and time periodic. It “locks in” to the forcing signal. The resulting
strong spanwise 2-D structures appear to further decrease the size of the separa-
tion bubble when compared with the PVGJ control.
The wall pressure coefficient as computed from the temporal and spanwise aver-
age of the simulation data and the attached flow pressure distribution obtained from
a design code (MISES) (Reimann et al., 2006) are shown in Fig. 45a. Both control
schemes delay flow separation and thereby increase the length of the favorable pres-
sure gradient region. The earlier suction side reattachment leads to a stronger pres-
sure recovery in the aft part of the blade. Wall normal velocity profiles are shown in
Fig. 22b. Slightly upstream of the separation location, at x/Cx = 0.5, the computed
velocity profiles are almost identical. At x/Cx = 0.7, all profiles indicate separated
flow, even for the cases with flow control. The thickness of the separation bubble is
about the same for both cases with AFC with a steeper gradient of the velocity
profile at the inflection point for the case with harmonic blowing through a slot. At
x/Cx = 0.9, the thickness of the separation bubble for the uncontrolled 3-D case is
still larger than at x/Cx = 0.7 while in the 2-D simulation the flow is already reattached
as a result of the overly strong spanwise flow structures. At this downstream location
attached flow profiles are also obtained for the cases with flow control. A larger wall
shear and smaller displacement thickness was found for the case with harmonic
blowing through a slot. Apparently, the spanwise coherent structures that are
300 H. F. FASEL ET AL.

Fig. 44 Spanwise vorticity (left) and streamfunction (right) from 3-D FSM (L1M).
Control by pulsed VGJs (top) and harmonic blowing and suction through slot (bottom)
(Reprint from Gross and Fasel, 2007b).

LIVE GRAPH
Click here to view
LIVE GRAPH
Click here to view

Fig. 45 a) Wall pressure coefficient and b) wall normal velocity profiles (L1M)
(Reprint from Gross and Fasel, 2007b).
TURBOMACHINERY APPLICATIONS 301

generated by this type of flow control facilitate a stronger wall normal momentum
exchange than the momentum exchange caused by turbulent mixing as seen for the
case with control by PVGJs. Separation control by PVGJs results in a 14% increase
in cy /cx and separation control by harmonic blowing results in an 18% increase in
cy /cx relative to the uncontrolled flow.

5. Closed-Loop Control (2-D)


Most of the AFC applications to date employ open-loop control. For open-loop
control, the controller parameters are optimized a priori for a given operating
point. Despite the proven effectiveness of open-loop control, it is not necessarily
robust and it works only for one operating point unless “look-up tables” are uti-
lized. These potential limitations of open-loop control can be remedied in closed-
loop flow control where the instantaneous flow data are collected, processed, and
fed back to actuators. The simplest form of closed-loop control is to feed back a
filtered downstream sensor signal to an upstream actuator (e.g., Israel et al., 2002;
Roussopoulos, 1993). Considerably more complex closed-loop flow control
schemes which may be based on models that describe the flow dynamics and the
flow response to the flow control may be conceived. Notwithstanding questions
concerning their practical realization, many other related issues such as control-
lability (for a certain actuator arrangement, how much authority over the flow
dynamics can the control scheme possibly have?) and observability (for a certain
set of sensors, how inclusive and accurate is the information about the flow state
and/or dynamics?) have to be considered as well. In the following, two simple
examples that show some of the characteristics of closed-loop control compared
to open-loop control are discussed.
A simple but robust and very efficient closed-loop controller was realized
(Gross and Fasel, 2005a) by feeding back data from a pressure sensor located at
90% axial chord, p0.9, to an upstream actuator

vjet = 10 ∂p0.9 /∂t (6)

For this controller (and for the slightly more elaborate controller discussed hereafter)
the sensor signal was low-pass filtered to remove high-frequency jitter. The separated
flow naturally amplifies spanwise disturbances into spanwise coherent structures
(shear layer instability of separated boundary layer and, possibly, global instability of
separation bubble) that facilitate flow reattachment. The dominant signal that a down-
stream pressure sensor will pick up is related to these structures. When this signal is
fed back to an upstream actuator, the most amplified structures will be forced and the
control effort will be small. As the flow is successfully controlled, the dimensions and
characteristics of the separation bubble change. As a consequence, the most ampli-
fied frequency also changes. This simple controller will always feed back the most
amplified frequency and thus respond to this change. The controller was tested in 2-D
simulations for the Pack B blade where the actuator was modeled as a slot.
The actuator signal quickly (within 4–5 shedding cycles) becomes strongly
time-periodic, indicated by a dominant peak in the frequency spectrum at fe  5
(Fig. 46) which is identical to the frequency that was found optimal for open-loop
control (Gross and Fasel, 2005b). The gain in cy /cx is 17% (open-loop: 17.6%).
302 H. F. FASEL ET AL.
LIVE GRAPH LIVE GRAPH
Click here to view Click here to view

Fig. 46 Actuator signal in the a) time-domain and b) frequency domain (Pack B,


S/Cx = 1.1) (Reprint from Gross and Fasel, 2005a).

For this simple controller, amplitude and phase-delay of the feedback may not be
optimal. Therefore, a slightly more sophisticated controller was developed (Gross
and Fasel, 2007c)

∂p
v jet = K P p (t - Dt P ) + K D (t - Dt D ) (7)
∂t

Using a gradient descent algorithm the parameters KP, KD, DtP, and DtD were
adjusted such that a control objective
2
t
1 Ê cx ˆ 1
J = Ú Á c ˜ + v2jet dt (8)
t =t -DT
2Ë y¯ 2

was minimized. The pressure sensor was located on the suction side at 81%
axial chord. The parameters KP, KD, DtP, and DtD were initialized with 0.5,
0.005, 0.02, and 0.02. The controller was tested for the L1M blade. The con-
troller parameters quickly converged to almost constant values (Fig. 47).
Dominant peaks in the frequency spectrum are located at fe  5.7 (amplitude
A  0.1) and fe  11.3 (amplitude A  0.06). The parameters of the first peak
are close to the optimum values found in the open-loop parameter study. The
gain in aerodynamic performance with respect to the uncontrolled case is 23%
(open-loop: 24%).

C. Approach 2: Direct Numerical Simulations of Active Flow Control


for Model Geometry
1. Simulation Details
In this approach, we investigated boundary layers on flat and curved plates sub-
jected to the same streamwise pressure gradient as measured in the experiments.
TURBOMACHINERY APPLICATIONS 303
LIVE GRAPH
Click here to view

Fig. 47 Time evolution of controller parameters and Fourier-transform of forcing


signal (L1M) (Reprint from Gross and Fasel, 2007c).

The displacement thickness Reynolds number at the separation location was the
same as in the experiments. The wall curvature for the curved plate matches that of
the suction side of the blade used in the experiments. The deliberate simplification
allowed us to focus all computational resources on the regions of primary interest
(the regions of actuation by the VGJs, boundary-layer separation and transition)
and to resolve all relevant scales from the laminar to the turbulent regime. Thus, the
relevant flow physics associated with LPT separation and transition as well as its
control using steady and pulsed VGJs could be investigated in all necessary detail.
For these very well resolved DNS we employed a highly efficient code based on
the incompressible Navier–Stokes equations in vorticity–velocity formulation
(Meitz and Fasel, 2000). The current version of the code employs fourth-order-
accurate compact finite differences in combination with a fourth-order-accurate
explicit Runge–Kutta time integration. The spanwise direction is treated with a
pseudospectral approach, which results in very high accuracy.
A schematic of the two computational setups (flat and curved plate) for the
incompressible code is shown in Fig. 48. A laminar boundary-layer profile is pre-
scribed at the inflow boundary. The separation bubble is generated by specifying an
appropriate boundary condition for the normal velocity component at the upper
boundary of the computational domain. This is equivalent to imposing a stream-
wise pressure gradient. Near the outflow boundary a buffer domain is employed
which acts like a sponge and dampens flow structures. The buffer domain prevents
unphysical disturbance reflections near the outflow boundary which is essential for
accurate unsteady simulations. Flow periodicity is enforced in the spanwise direc-
tion. Further details can be found in Postl et al. (2004) and Postl (2005).

2. Uncontrolled Flow
Typical results from simulations of the uncontrolled, separated boundary layer
are shown in Figs. 49 and 50. For the flat plate the boundary layer separates (in the
mean) at a suction surface length* (SSL), s, of approximately s = 5.7 and reattaches
at approximately s = 16.8. The SSL was measured from the LE of the blade. Note
that the TE of the hypothetical LPT suction surface is located at Cs = 10.18,
304 H. F. FASEL ET AL.

Fig. 48 Schematic of computational domain for a) flat-plate and b) curved plate


cases (Reprint from Balzer et al., 2007).

indicating that the flow for an actual LPT cascade would, for this case, not reattach
to the blade surface. The separation and reattachment locations for the uncontrolled
flow over the curved plate are at approximately s = 4.7 and s = 9.1, respectively.
The wall pressure coefficient vs SSL for the uncontrolled case is shown in
Fig. 51a in comparison to the experiments. The pressure plateau downstream of
62% SSL indicates the region of boundary-layer separation. Velocity profiles at
various streamwise locations are also shown in Fig. 51b. The experimental profiles
LIVE GRAPH
Click here to view

Fig. 49 Uncontrolled boundary layer flow over flat plate model geometry. a)
Instantaneous contours of spanwise vorticity; b) time- and spanwise-averaged stream-
lines. Dotted line shows location of hypothetical TE (Reprint from Postl, 2005).

*The SSL was non-dimensionalized by 1 in, where the axial chord length of the blade was C = 7 in.
x
The given length scales have to be divided by 7 to obtain a non-dimensionalization by the axial chord
length.
TURBOMACHINERY APPLICATIONS 305
LIVE GRAPH LIVE GRAPH
Click here to view Click here to view

Fig. 50 Uncontrolled flow over curved plate model geometry. a) Time- and spanwise-
averaged streamlines; b) instantaneous contours of spanwise vorticity (Reprint from
Balzer et al., 2007).

do not show negative velocities because they were obtained from hot-wire mea-
surements. While the experimental profile at 92% Cx indicates attached flow, it is
in fact separated. Overall, the careful setup of the numerical simulations resulted
in a good agreement with the experimental data for the base flow.

LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view
Fig. 51 Uncontrolled boundary-layer flow over model geometry: a) Wall pressure
coefficient; b) streamwise velocity profiles. Comparison between simulations and mea-
surements by Bons et al. (2001a), Huang et al. (2006a) and Sondergaard et al. (2002a)
(Reprint from Balzer et al., 2007).
306 H. F. FASEL ET AL.

3. Controlled Flow
Two different VGJ configurations were investigated: jets issued vertically into
the boundary layer and jets that were pitched (30 deg) and skewed (90 deg) to the
free-stream direction (Fig. 4). For both configurations, steady and pulsed actua-
tion was considered. The VGJ holes were spaced at 0.135Cx (0.056Cx in the
experiments by Sondergaard et al., 2002a), had a diameter of 0.015Cx (0.0056Cx in
the experiments by Sondergaard et al., 2002a), and were resolved with 24 × 25
points. A cos3 velocity distribution was prescribed over the jet exit holes. Other
combinations of hole spacing and hole diameter that are closer to the experimental
parameters are possible. Our choice is a compromise where we tried to get close
to the geometric VGJ parameters used in the experiments while keeping the com-
putational expense at a reasonable level.
a. Steady VGJs. We first investigated steady VGJs. For a blowing ratio of
B = 0.316 (cm = 1.2 × 10-4), vertical vs angled steady VGJ injection is compared in
Fig. 52. Flow structures are identified using the l2 vortex criterion by Jeong and
Hussain (1995). Following a rapid decay immediately downstream of the forcing
location, the longitudinal vortices generated by the jets become amplified in the
region of strong streamline curvature associated with the separating boundary
layer. A local stability analysis of streamwise velocity profiles indicated that a
Görtler instability mechanism (due to streamline curvature caused by the sepa-
rated boundary layer) may be responsible for this amplification.
A detailed investigation of the unsteady flow structures reveals that the final
stages of the laminar-turbulent transition process in each of the two cases (angled
and vertical jets) can be characterized by the formation of hairpin-like vortices
(Fig. 53). The breakdown to turbulence occurs more rapidly for angled injection
which can be attributed in part to the (in the spanwise direction) deeper penetra-
tion of the jets into the boundary layer. This is illustrated in Fig. 54. However,
from this and many other simulations we find that, once the flow is separated, an
accelerated breakdown to turbulence by itself does not necessarily provide the
most effective mechanism for optimal separation control. In other words, the mix-
ing associated with the small-scale flow structures appeared to be rather weak, and

Fig. 52 Flow control using steady VGJs with blowing ratio B = 0.316. Time-averaged
isosurfaces of l 2 = -2 and isocontours of spanwise vorticity wz. Lines indicate loca-
tions of w z = 0 (separation line). a) vertical injection and b) angled injection (Reprint
from Postl, 2005).
TURBOMACHINERY APPLICATIONS 307

Fig. 53 Flow control using steady VGJs with blowing ratio B = 0.316. Instantaneous
visualizations (top and side views) of iso-surfaces of l 2 = -50. a) Vertical injection; b)
angled injection. Dotted lines show spanwise averaged reattachment point (Reprint
from Postl, 2005).
LIVE GRAPH
Click here to view

Fig. 54 Flow control using steady VGJs with blowing ratio B = 0.316. a) Time-
averaged streamlines emanating from the center of the VGJ holes and b) iso-contours
of streamwise vorticity at s = 7 (Reprint from Postl, 2005).
308 H. F. FASEL ET AL.

Fig. 55 Flow control using steady VGJs with blowing ratio B = 2. Instantaneous visu-
alizations (top views) of iso-surfaces of l 2 = -50. a) Vertical jet injection and b) angled
jet injection (dark shaded areas indicate separated flow) (Reprint from Postl, 2005).

a considerable streamwise distance was required before enough free-stream


momentum was entrained to reattach the flow.
To understand the effect of the blowing ratio, we performed simulations where
we increased the blowing ratio up to B = 2 (cm = 4.7 × 10-3). Mean flow results
indicate that, up to a “threshold” blowing ratio of B = 1, angled VGJs were more
effective than vertical VGJs in reducing flow separation. However, from many
other simulations (not shown here) we learned that beyond this threshold value the
trend was reversed and vertical jet injection was more effective, resulting in fully
attached flow along almost the entire surface (in the spanwise average). A com-
parison for the two cases (vertical vs angled) by instantaneous visualizations of
isosurfaces of l2 = -50 for a blowing ratio of B = 2 is shown in Fig. 55.
Comparison with Fig. 53 confirms that the dominant physical mechanism has
changed as the jet amplitude was increased. For B = 2, with vertical injection, a
strong horseshoe vortex develops that “wraps around” the column of fluid
injected by the jet (Fig. 56). As a result of this horseshoe structure, entrainment
of high-momentum fluid from the free-stream is increased significantly, thereby
effectively generating a “new” boundary layer behind the jet exit hole. Due to
the lack of symmetry associated with the cross-stream injection of the angled
VGJs, the oncoming boundary layer for the case with angled VGJs was found to
be mainly diverted in the direction of the jet injection. Consequently, only “one-
legged” horseshoe structures developed, thus leaving one side of the surface
essentially unaffected by the forcing. In fact, as a result of the entrainment of
high-momentum fluid on the upper side of the surface (when viewed from
above), low-momentum fluid appeared to be transported to the lower side,
thereby further destabilizing the flow which is already at the verge of separation.
This observation led to the conclusion that, for angled VGJs, a further increase
of the blowing ratio may not lead to an attached flow along the entire surface
(for the present hole spacing and diameter). Comparison of Figs. 53 and 55
also illustrates that the laminar-turbulent transition process was delayed at the
larger blowing ratio and that the primary control mechanism for both cases was
laminar in nature. Also, and this is consistent with experimental observations
TURBOMACHINERY APPLICATIONS 309

Fig. 56 Visualization of the horseshoe vortex. Time-averaged streamlines passing


through a horizontal rake located at s = 4.8, y/d  0.14, and jet-exit velocity vectors.
(Reprint from Postl, 2005).

(Hansen and Bons, 2006), the streamwise vortices maintain their coherence
over a longer downstream distance for angled injection.
b. Pulsed VGJs. We then investigated pulsed VGJs. First, results are
presented for angled and vertical jet injection for a forcing frequency of F+ = 5.4
(normalized with axial chord and inflow velocity), a duty cycle of t = 10%, and a
blowing ratio of B = 1, which results in cm = 1.2 × 10-4 (computed with axial chord
times VGJ spacing as reference area). Since the flow structures that are generated
downstream of the forcing location (Fig. 57a) closely resemble the hairpin vorti-
ces that develop in the late stages of Klebanoff-type transition scenarios (see
Fig. 57b), a “bypass” transition mechanism was at first considered to be the most

Fig. 57 a) Instantaneous visualization (l2 = -5) of flow control using pulsed vertical
VGJs (t = 10%, B = 1, F+ = 5.4) for flat plate model geometry (Reprint from Postl,
2005). b) L-vortices in a transitional flat plate boundary layer (Reprint from Bake
et al., 2002).
310 H. F. FASEL ET AL.

Fig. 58 Flow control using pulsed vertical VGJs (t = 10%, B = 1, F+ = 5.4).


Instantaneous flow visualizations (iso-surfaces of l2 = -2
-25); top and side view (Reprint
from Postl, 2005).

relevant mechanism for controlling LPT separation with pulsed VGJs. However,
from other simulations we realized that other mechanisms may also play signifi-
cant roles. An indication for this is already seen in Fig. 57a as the hairpin vortices
and longitudinal structures that were generated by the pulsed VGJs decay in the
streamwise direction. This led us to speculate that the bypass mechanism and the
resulting vortical structures may only be indirectly responsible for the effectiveness
of pulsed VGJs.
More important, and surprising at first, the simulations revealed that, for both
vertical and angled injection strong spanwise coherent vortical structures were
found to develop in the separated flow region and are followed further downstream
by the rapid generation of smaller and smaller scales (Fig. 58). Of particular
importance was the observation that the developing small-scale structures were
part of large-scale, spanwise coherent structures. This observation led to the con-
jecture that the formation of these structures may be the primary cause for the
increased effectiveness of pulsed VGJ actuation vs steady VGJs.
A proper orthogonal decomposition (POD) (Lumley, 1967; Sirovich, 1987) of
the time-dependent flow data confirmed this conjecture. Results of the POD analy-
sis for the case of vertical injection are shown in Fig. 59. The spanwise coherent
structures are indeed by far the most energetic unsteady flow structures for both

Fig. 59 Flow control using pulsed vertical VGJs (t = 10%, B = 1, F+ = 5.4). a) POD
mode 1 for the case with vertical injection and b) normalized forcing signal at the VGJ
injection location and POD time functions associated with modes 1 and 2 (Reprint
from Postl, 2005).
TURBOMACHINERY APPLICATIONS 311

vertical and angled pulsed injection. The formation of the structures is in perfect
phase with the pulsed actuation in both instances indicating that these structures
are generated by the pulsed VGJs. It is obvious that the dominant physical mecha-
nism must somehow be similar for both the vertical and angled jet injection. This,
of course, is contrary to the findings for steady jets, as discussed previously.
We conjectured that the similar effectiveness of angled and vertical jets must be
due to the unsteady (periodic) pulsing and that resulting large-scale spanwise
coherent structures must, therefore, be a consequence of the 2-D disturbance
component of the localized forcing (as a result of employing the jets “in phase”).
Due to the shear layer instability of the separated boundary layer (inflection point
in the velocity profiles in the separated region), the 2-D disturbances undergo a
strong amplification in a downstream direction. This was confirmed by plotting
the amplitude of the 2-D component of the u' velocity disturbance vs the down-
stream direction (Fig. 60). After a transient, downstream of the exit holes for the
VGJs (the jets are located at s = 5), the 2-D disturbances experience approximately
exponential (in the log-plot linear) growth. Exponential growth is an indication of
a linear instability mechanism. Thus, theu' main conclusion here is that the control
is so effective because a linear instability mechanism is exploited. As a conse-
quence of this mechanism, the energy required for the amplification of the 2-D
disturbances is provided by the base flow and comes “free of charge”, thus requir-
ing only very small actuator amplitudes (and thus very small cm).
Additional results of simulations using pulsed VGJs at various lower or higher
frequencies (relative to the “baseline case” frequency, F+ = 5.4) are shown in
Fig. 61. For these simulations duty cycle (t = 10%) and blowing ratio (B = 1)
were held constant. Forcing with lower forcing frequencies had relatively little
effect on the extent of the separated flow region when compared to the baseline
case, while increasing the frequency lead to a significant reduction in the control

LIVE GRAPH
Click here to view

Fig. 60 Fourier amplitude of the 2-D component of the u¢ disturbance velocity (max
over y) for the fundamental frequency F+ = 5.4. Open symbols: vertical injection;
closed symbols: angled injection (Reprint from Postl, 2005).
312 H. F. FASEL ET AL.

LIVE GRAPH
Click here to view

Fig. 61 Flow control using pulsed vertical VGJs for various pulsing frequencies
(t = 10%, B = 1). a) Contours of spanwise wall vorticity w z (dark: separated flow; light:
attached flow) and b) frequency spectra for pulsed VGJs (Reprint from Postl, 2005).

effectiveness. This can be explained by the hydrodynamic stability characteris-


tics of the separated shear layer. Thus, the results of Fig. 61 are an additional
confirmation that for most effective and efficient control the instability of the
underlying flow needs to be exploited. As shown by the amplitude spectra in
Fig. 61b, pulsed forcing with a fixed duty cycle introduces a number of higher
harmonics of the fundamental pulsing frequency, and the spectrum shifts to the
right or left depending on the value of the fundamental pulsing frequency.
Consequently, as long as the pulsed forcing generates frequencies to which the
flow is hydrodynamically unstable, the amplification of the resulting instability
modes will yield an effective control of the separation. For a more detailed dis-
cussion of this concept the reader is referred to Chapters 2 and 4.
We also carried out additional simulations where we investigated the effect of
the duty cycle in the range from 10–100% (100% means steady forcing) and
adjusted the blowing ratio for keeping the jet momentum coefficient cm constant.
In all cases control was almost equally effective except, when t approached 100%
(steady VGJs), the separation length increased significantly (not shown). Although
this observation alone was not surprising, the observation that this increase hap-
pened very “suddenly” certainly was. This finding further corroborated the con-
jecture that the exploitation of an inviscid linear hydrodynamic instability
mechanism was primarily responsible for the stunning effectiveness of pulsed
VGJs for LPT separation control. The observation that all the unsteady cases were
almost equally effective led to the conclusion that, from an engineering point of
view, the choice of the duty cycle does not seem to be as critical as that of the
pulsing frequency, again a confirmation of the underlying 2-D instability mecha-
nism. Lower-duty cycles are preferable as they require a smaller mass flux (when
the momentum coefficient cm, is not kept constant). Instantaneous flow visualiza-
tions (Fig. 62) and results obtained from a POD analysis of the flow data (Fig. 63)
reveal that, for intermediate duty cycles (here, t = 40%), flow structures are being
concentrated in areas with oblique coherence. This stunning observation leads to
TURBOMACHINERY APPLICATIONS 313

Fig. 62 Flow control using pulsed vertical VGJs. Instantaneous visualizations (top
views) of isosurfaces of l 2 = -2. a) Duty cycle t = 10% and b) t = 40% (Reprint from
Postl, 2005).

the conjecture that the amplification of 3-D instability modes may lead to the
formation of oblique coherent structures.
c. Pulsed VGJs: Curved Plate Geometry. Results for the curved wall
geometry confirmed our findings for the flat plate geometry. Simulations with
pulsed vertical VGJs and identical control parameters (t = 10%, B = 1, F+ = 5.4)
showed a similar effectiveness of the flow control (Fig. 64). Again, compared to the
uncontrolled flow, pulsed VGJs were found to result in an earlier transitioning and
the appearance of spanwise coherent structures that were amplified in the stream-
wise direction. Compared to the flat plate, we noticed a higher amplification of the
spanwise structures between s = 7 and s = 8 (Fig. 65) but also an earlier loss of
spanwise coherence of these structures in the downstream direction due to the
stronger amplification of 3-D modes. The striking similarity between the flat and

Fig. 63 Flow control using pulsed vertical VGJs. POD mode 1, top view towards the
surface. a) Duty cycle t = 10% and b) t = 40% (Reprint from Postl, 2005).
314 H. F. FASEL ET AL.

Fig. 64 Curved plate geometry. Flow control using pulsed vertical VGJs (t = 10%,
B = 1, F+ = 5.4). a) Instantaneous flow visualization (isosurfaces of l 2 = -25) and
b) POD mode 1 (Reprint from Balzer et al., 2007).

LIVE GRAPH
Click here to view

Fig. 65 Fourier amplitude of the 2-D component of u' disturbance velocity (max over
y) for the fundamental frequency F+ = 5.4. Comparison of curved and flat plate geom-
etry results (Reprint from Balzer et al., 2007).
TURBOMACHINERY APPLICATIONS 315

Fig. 66 Curved plate geometry. Flow control using pulsed vertical VGJs. POD mode 1,
top view towards the surface. a) Duty cycle t = 10% and b) t = 40%.

curved plate results can be explained by the low curvature of the Pack B LPT blade
near the TE which is where flow separation occurs for the uncontrolled flow.
As for the flat plate geometry, a variation of the duty cycle also led to the
appearance of oblique coherent structures (Fig. 66). However, these structures
were not as pronounced as for the flat plate (Fig. 63) indicating that this mecha-
nism may be less relevant for LPT applications.

IV. Discussion
A. General Comments
The relevant physics associated with AFC for LPT flows are highly complex, as
both unsteady separation and transition mechanisms are at work interactively.
Each of these areas alone, transition from laminar to turbulent flow and unsteady
separation, belong to the least understood areas of flow physics. The main under-
standing of transition is based on the so-called linear regime, where the amplitudes
of the instability waves are small and, as a consequence, the so-called linear sta-
bility theory can be employed for modeling. For LPT applications, as a conse-
quence of the hydrodynamic instability of the separating boundary layer,
disturbances quickly reach very large amplitudes. The strong amplification is due
to the streamwise adverse pressure gradients, the convex wall curvature (suction
side of blade), and of course the inviscid instability mechanism when the bound-
ary layer is separating. Therefore, primary linear theory is no longer applicable
farther downstream and in particular cannot capture breakdown mechanisms to
turbulence. For high FST the transition process is of a bypass nature (Morkovin,
1969), that is the linear stages are partially or completely bypassed. Due to the
non-linearity and the non-uniqueness of the bypass mechanisms, slight changes in
316 H. F. FASEL ET AL.

“initial” conditions (operating conditions) can result in drastically different


breakdown-to-turbulence scenarios.
The fundamental understanding of separation is almost as incomplete as that of
transition, especially when the separation process is unsteady and 3-D as in LPT
flows. This unsteadiness is introduced by the pulsing of the VGJs and/or by the
naturally present large coherent flow structures, which result from the instability
of the separated base flow (or in practical applications by the wakes shed from the
upstream stages). The three-dimensionality is caused by the fact that the jets are
injected through small holes that are relatively far apart from one another (several
hole diameters). Again, as for transition, boundary-layer separation under such
conditions is a highly non-linear, non-unique process that exhibits a strong sensi-
tivity to initial conditions.
Clearly, for LPT separation, the two mechanisms interact non-linearly, thereby
considerably expanding the range of non-uniqueness. Separation, in general,
strongly accelerates transition while transition, in general, delays or can even
prevent separation. However, to what degree they affect each other depends on the
details of the initial conditions and is strongly influenced by the geometry (rough-
ness, wall curvature, wall temperature, jet geometry, frequency, and amplitudes of
forcing, etc.). Thus, when both of these non-linear, non-unique mechanisms are at
work at the same time, as is the case for the LPT, surprises are likely, both positive
and negative. For example, when frequencies and amplitudes of the pulsed blow-
ing are “just right”, the effectiveness of separation control is indeed stunning,
requiring a very small energy input. In other instances, although unintentional and
often due to a lack of understanding, AFC is not effective in the sense that separa-
tion is not prevented/delayed or that an unacceptable energy input is required.
Therefore, in light of the complex physics that are at work in AFC for LPTs using
VGJs, it is obvious that a better understanding of the most relevant physical
mechanisms needs to be achieved before this technology can be transitioned suc-
cessfully into practice and to ensure safe, reliable, and effective operation.

B. Physical Mechanisms Relevant for an Effective Separation Control


The synergisms between experiments (Bons et al., 2001a, 2002; Huang et al.
2006a, b; Sondergaard et al., 2002a, b), numerical simulations (Gross and Fasel,
2005a, b, 2007a; Postl, 2005; Postl et al., 2003, 2004; Rizzetta and Visbal, 2004,
2005, 2007), and linear stability analysis (Abdessemed et al., 2004, 2006; Theofilis
and Sherwin, 2004) resulted in a significant step forward in the understanding of
the physical mechanisms that are relevant for LPT separation control. The experi-
ments at AFRL by Rivir and co-workers have convincingly demonstrated the
potential benefits of AFC using VGJs. It was conjectured that PVGJs resulted in
an accelerated breakdown to turbulence resulting in larger mixing and free-stream
entrainment, thereby suppressing separation. This was confirmed by CFD results
by Rizzetta and Visbal (2004, 2005), Postl (2005), and Gross and Fasel (2007a).
Although local and global instabilities cannot be separated in CFD simulations,
simulations of the uncontrolled “natural” flow which displayed vortex shedding
suggest that the separation bubble is absolutely unstable (Gross and Fasel, 2005a, b,
2007a; Rizzetta and Visbal, 2003a, 2004, 2005). Two-dimensional simulations
with forcing through a slot where separation is strongly suppressed for small
TURBOMACHINERY APPLICATIONS 317

forcing amplitudes indicate that flow instabilities result in an amplification of the


control input and can be exploited, resulting in a very efficient flow control.
However, for this amplification to occur, the disturbances must be introduced in
the right location and manner (receptivity) and at the right frequency and ampli-
tude. For pulsed VGJs 3-D disturbances are introduced that result in bypass type
transition (Gross and Fasel, 2007a; Postl, 2005; Rizzetta and Visbal, 2005).
In-phase actuation, however, also introduces a 2-D disturbance component, which
is amplified by the flow and leads to the formation of spanwise coherent structures
which further assist in keeping the flow attached. Because of the amplification of
the controlled disturbance input, because of the exploitation of a linear hydrody-
namic instability mechanism, an effective flow control becomes possible for rela-
tively small actuation amplitudes. This is in contrast to when VGJs are operated in
a steady fashion for generating streamwise vortices (Hansen and Bons, 2006).
Steady VGJs require much larger cm (and thus energy) for effective control because
no linear (primary) hydrodynamic instability mechanisms can be exploited.

C. Cost/Gain Estimate
The following analysis was carried out based on experimental results by Bons,
Sondergaard, Rivir and co-workers for the Pack B blade with a design blade spac-
ing of S = 0.88Cx and for Re = 25,000 and a FSTI of 1%. Sondergaard et al.
(2002b) state that with steady VGJs for B = 2 and for their specific VGJ configura-
tion (VGJ hole diameter/chord length = 1 mm/8.9 cm, hole spacing identical to
10 times the hole diameter) the mass flux through the VGJ holes was approxi-
mately 0.2% of the total mass flux through the cascade at the design blade spacing.
For pulsed VGJs, the corresponding mass flux can be obtained by multiplying this
number by the duty cycle t, which can be as low as 1% (Fig. 13b), resulting in a
mass flow ratio of 0.01 × 0.2% = 0.002%.
The wake loss coefficient g, for the uncontrolled flow was approximately 0.45.
With both steady and pulsed VGJs the loss coefficient could be reduced to about
0.18 (Bons et al., 2002; Sondergaard et al., 2002a). The blowing ratio B required
for an effective flow control with steady VGJs was B = 1.5 (Fig. 8). For pulsed
VGJ actuation with B = 2 and t = 1%, separation could also be successfully con-
trolled (Fig. 11). The momentum coefficient for their specific VGJ configuration
was cm  0.00121B2t where the area was normalized by 72% of the axial chord
length Cx (the approximate length of the separated flow region for the uncontrolled
flow) and the span and the blowing ratio B, was based on the local flow velocity,
vlocal at the VGJ location (which was roughly two times the cascade inlet velocity
vin). Dividing the momentum coefficient by the change in the wake loss coefficient
Dg, and taking into account that the wake loss coefficient is integrated over the
wake S, an estimate for the VGJ momentum input required for a certain change in
wake momentum is obtained, 0.5(vlocal /vin)2[0.72(Cx /S)(cm /Dg ) = 1.64](cm/Dg ).
With Dg = 0.45 – 0.18 = 0.27, for steady blowing with B = 1.5 (cm = 0.0027) this
momentum ratio is 1.6% and for pulsed actuation with B = 2 and t = 1%
(cm = 4.84 × 10-5) a value of 0.03% is obtained.
In the Pack B experiments by Sondergaard et al. (2002a) at the design blade
spacing of 0.88Cx, the lift coefficient was found to increase by about DcL = 0.18
when separation was eliminated with AFC. The power required for the VGJ
318 H. F. FASEL ET AL.

actuation is _12 v3A, where v is the jet exit velocity and A is the total area of all VGJ
holes. An estimate for the power extracted per unit span from the cascade is
_1 rv2 C c v
2 in x L rotor assuming that the cascade blades rotate at a constant wheel speed.
The rotor wheel speed is here assumed to be vin sin 55 deg/f, where f is the flow
coefficient (typically f = 0.8). Taking the ratio of the power expense to the change
in extracted power yields a value of 6% for steady VGJ actuation and 0.15% for
pulsed actuation.
A final benefit of the use of VGJs is derived from the reduced total pressure loss
in the cascade wake. Since total pressure loss represents lost work (Denton, 1993),
any reduction in total pressure loss results in additional work that can be extracted
in subsequent turbine stages. For the same data set cited above, the power required
for VGJ actuation represents roughly 6% (for steady actuation) and 0.15% (for
pulsed actuation) of the power gained through total pressure loss reduction in the
cascade wake. Since this benefit is in addition to the increased work extracted
from the higher blade cL, the combined cost is only 3% for steady VGJ actuation
and 0.07% for pulsed VGJ actuation.
There are additional losses that are not included in this analysis. Because the
mass flow required for the VGJ actuation is extracted from the compressor, there
is lost work because the VGJ mass flow was compressed but not combusted. Then
there are additional total pressure losses in the ducting from the compressor to the
VGJs which are difficult to estimate. Nevertheless, for the experimental condi-
tions considered here, AFC with steady and pulsed VGJs is clearly worth the
effort. Pulsed actuation results in a significant increase in effectiveness over steady
actuation. It has to be kept in mind that the performance loss in the uncontrolled
case can be lower in the real turbine environment as a result of FST and unsteady
wakes which reduce separation (Bloxham et al., 2007; Sondergaard et al., 2002a).
However, even then a significant cost benefit can be expected from AFC, in par-
ticular with pulsed VGJs as the gain-to-cost ratio for such actuators is very large.
Also, it was found that the blade spacing could be reduced by a factor of two
without significant performance losses by applying flow control with steady VGJs
and a blowing ratio of only four (Sondergaard et al., 2002b). This reduction in the
cascade solidity results in significant weight savings that are, however, diminished
by the additional weight of the components required to operate the VGJs.

D. Open-Loop vs Closed-Loop Control


The robustness of open-loop control schemes for LPT separation has not been
investigated. A control strategy that is optimized for one design operating point is
certainly not optimal for an off-design operating point. It may, however, still per-
form satisfactorily (in particular when “look-up tables” are used) and thus not
warrant the extra expense related to more complicated closed-loop control
schemes. Although closed-loop control may be required for other AFC applica-
tions such as the control of cavity tones, their immediate benefit for LPT separa-
tion control has to be proven.
The main advantages of closed-loop control are its inherent robustness and capa-
bility to respond to changing operating conditions. In this chapter we presented two
simple closed-loop control schemes that exploited the fact that the flow amplifies
disturbances. This was shown to be beneficial for LPT separation control since it
TURBOMACHINERY APPLICATIONS 319

made the control more effective. The design of the control scheme was motivated by
physical reasoning. In other instances, however, it may be desirable to suppress flow
structures, or it may not be directly obvious how an effective closed-loop control
could be realized. In such instances, and whenever more sophisticated closed-loop
controllers are desired, a common strategy is to derive reduced-order models of the
flow and employ these models for controller development. Such models typically
consist of a limited number of coupled equations for describing the dynamics of the
flow and its response to AFC near one or more operating points. The fact that such
models are valid only for a limited number of operating points is both a blessing and
a curse, as without this assumption the derivation of reduced-order models would be
impossible and because this assumption limits the applicability of the models. In
summary, despite the many difficulties associated with closed-loop control, it is a
very attractive alternative and offers potential gains over open-loop control. The
development of effective model-based closed-loop control schemes for LPT appli-
cations is the subject of ongoing research and is likely to lead to a number of suc-
cessful closed-loop control applications.

E. Challenges/Difficulties for Future Implementation of AFC in a


Jet Engine
AFC by pulsed VGJs may have great potential for significant performance
increases of future engines. However, there are major challenges for technical
realization, and the increased complexity will increase the cost associated with
production and maintenance. The manufacturing techniques required for drilling
VGJ holes in the blade exist (similar techniques are being used for fabricating the
holes required for shower head cooling in the high-pressure turbine). High-
pressure air could be bled from the compressor stage. This air would have to be
modulated using solenoid valves or other simpler mechanical devices such as a
perforated rotating disk that is driven by the engine. The implementation of plasma
actuators appears more difficult as these devices have a short lifespan (demanding
shorter LPT maintenance intervals) and cause high-frequency radio interference.
Before any of these technologies are implemented additional studies are required
for determining how effectively and reliably such AFC devices work in the harsh
turbine environment which is characterized by flow three-dimensionality and
rotational forces, vibration, free-stream turbulence, wakes, and significant surface
quality degradation.

F. Lessons Learned from LPT Case: Transition of Technology


to Other Applications
One may consider transitioning the knowledge and experience gained for the
LPT to other turbine engine components. Operating Reynolds numbers for high-
pressure turbine (HPT) and compressor stages are usually considerably higher
than for LPT stages. The control schemes that were identified to be most effective
for LPT separation control (pulsed VGJs and pulsed blowing through slot/plasma
actuators) exploit hydrodynamic instabilities that result either in an accelerated
transition or a transition delay. Both phenomena are less relevant at high Reynolds
number conditions. Here, the flow can be expected to transition immediately after
320 H. F. FASEL ET AL.

separation. In addition, for the HPT, very high levels of FSTI are encountered.
Flow control schemes that exploit possible instabilities of the turbulent flow which
manifest themselves in the form of coherent structures may be adequate for such
flows. Heat transfer is another concern, especially for the HPT. Increased wall
normal mixing is undesirable as it increases the heat load. The flow control
schemes discussed for the LPT control which reduce flow separation by increas-
ing the entrainment of free-stream fluid may, therefore, be inappropriate for HPT
applications.
The technology may, however, be transitioned to other applications where lami-
nar separation has to be avoided in off-design conditions, such as laminar airfoils
at high angles of attack, or in low Reynolds number aerodynamic applications
such as small UAVs. It may also be employed to replace the control surfaces of
small UAVs.

Acknowledgments
This work was performed under sponsorship from the Air Force Office of
Scientific Research, with T. Beutner and R. Jefferies as contract monitors. HPC
resources were provided by the DoD High Performance Computing Moderniza-
tion Program.
Chapter 10

Combustion Control

Suresh Menon* and Ben T. Zinn†


Georgia Institute of Technology, Atlanta, Georgia

I. Introduction
Combustion control may be one of the first active control approaches developed
at the beginning of human civilization. Starting a fire and/or keeping a fire lit (regard-
less of rain, wind, day or night) were essential for human species survival and perhaps
even played a fundamental role in creating an intelligent species. It is estimated that
fire was first controlled over 230,000 years ago (Goudsblom, 1986). Even from the
beginning, control of lean burning flames and avoidance of flame blowout were
perhaps the motivation for many ad hoc innovative active control techniques that
were developed. It is interesting to note that both lean blowout (LBO) and combus-
tion instability (CI) can result in or be a consequence of flame loss and/or structural
failure of the combustion source device, and they remain the two major vexing
problems to the present day. Although LBO can be easily identified as an inability of
the flame to sustain itself as the fuel flow or content is reduced, CI is much more
complicated since it may or may not be a result of lean burning. Regardless, in both
cases the combustion process is interrupted either temporarily or permanently. A
distinguishing feature that separates CI from LBO is that in some cases, CI results in
not only flame blowout but also serious damage to the system (Dowling, 2000;
Lieuwen and McManus, 2002; Lieuwen and Yang, 2005). Both LBO and CI are
tough problems to resolve because of the underlying nonlinearity of the physics. As
combustion system complexity and demands for low-emission, high-efficiency
systems continue to increase, the need to achieve successful, possibly “smart” con-
trol over a wide range of operating conditions is becoming more urgent.
The primary objective of an ideal combustion system is to use minimal fuel with
no complications, without any adverse affect to human society and/or environment,

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Professor. Associate Fellow AIAA.
†Regent’s Professor, David Lewis Chair. Fellow AIAA.

321
322 S. MENON AND B. T. ZINN

and still deliver whatever is demanded of such a system. This is not easily achieved
and, in the past, most power generation and propulsion systems were built to
deliver the necessary energy output with little regard to the amount of pollutant
emission (e.g., Lefebvre, 1995, 1999; Mellor, 1976). For example, emissions from
past and most current energy production systems (Gupta and Lilley, 1994;
Lefebvre, 1999) are now considered quite high, and with the increase in cost and
awareness of the impact of combustion on the environment, more efficient and
low-emission systems are being considered (Gupta, 1997; Koff, 1994; Mongia
et al., 2003; Richards et al., 2001; Walsh and Fletcher, 2004). This goal still
remains to be met but there are many encouraging results to report. Some recent
results are highlighted in this chapter.
Control can be both passive (i.e., the system hardware is permanently changed
to allow stable operation over the regime of interest) and/or active (i.e., the control
authority is implemented based on real-time observations). Within the active
control strategy, both open loop and closed loop systems are being developed and
evaluated. In general, both passive control techniques such as the use of baffles
and/or low-speed recirculation regions (as in dump combustors), and active con-
trol techniques using secondary air or fuel injection, acoustic forcing and/or
dynamic shape changes are not new but refinement of techniques developed thou-
sands of years ago during human control of fire. As in early times, human intuition
and experience still remain some of the major motive forces used for successful
control of combustion systems.
There is a considerable amount of work being done in active control for many
combustion applications (e.g., the extensive summary in Dowling and Morgans,
2005). Furthermore, there are many flavors of active control being developed and
demonstrated in both laboratory-scaled and full-scale devices. The differences in
many cases are minor, e.g., differences in the specific sensing and actuating
systems, but in most cases they employ similar strategies for control. On the other
hand, for LBO or CI control, strategies can be diverse since the phenomena being
controlled are different. It is also worth noting that there are subtle but significant
differences between sensing and avoiding LBO and/or CI before they occur (for
example, by using precursor signal information), and sensing and controlling LBI
and/or CI after they manifest themselves. Although an attempt will be made to
provide up-to-date references to all such studies it is not possible to do complete
justice to all approaches and results reported in open literature. Here, we sum-
marize some key observations from past studies and focus the discussion on active
flow and combustion control for combustion applications.
In this chapter, we will focus primarily on LBO- and CI-related active control
studies of systems used in planetary flight and/or power applications, i.e., gas
turbines (both gas and liquid fueled). The majority of combustion control investi-
gations in the literature are experimental and will dominate the discussion in this
chapter; however, available numerical and theoretical results will be introduced as
appropriate to each application.
Combustion control in operational devices requires a comprehensive integra-
tion of technologies from various disciplines, and thus demonstrates the practical
use of the state of the art in many technical fields. In particular, advances in
computer software and hardware, sensors, diagnostics, actuators, and the under-
standing of fundamental processes all need to come together in order to devise
COMBUSTION CONTROL 323

and implement a robust active control system (ACS) for combustion control.
Many of the underlying issues and mechanisms of combustion instabilities have
been discussed extensively in many publications (e.g., Candel et al., 1993; Ducruix
et al., 2003, 2005; Paschereit et al., 1999; Poinsot et al., 1987) and in books with
extensive references (e.g., Lieuwen and Yang, 2005; Poinsot and Veynante, 2005),
and therefore will not be repeated in detail here. There have been many successful
demonstrations of active control of combustion in both laboratory scale and opera-
tional (sector-scale) gas turbine combustors. These successes (and related obser-
vations) have been reviewed in many papers and or books (e.g., Candel, 1992,
2002; Candel and Poinsot, 1987; Docquier and Candel, 2002; Dowling and
Morgans, 2005; Lieuwen and Yang, 2005; Lieuwen and Zinn, 1998; McManus
et al., 1993; Zinn and Neumeier, 1997). Again, for brevity not all of these results
are discussed here.
The eventual objective of all these efforts is to develop and demonstrate an
“intelligent” engine that employs nonlinear, auto-adaptive control techniques for
continuous performance optimization and automatic reconfiguration to provide
condition-based performance (i.e., automatically accommodate for deterioration
and damage, to deliver the best possible performance even in a degraded state).
Some of these issues are discussed in cited references (e.g., Dowling and Morgans,
2005; Lieuwen and Yang, 2005). Clearly, this goal cannot be achieved by combus-
tion control alone and will require an integrated “intelligent” system that senses
all aspects of flow processes in the inlet, compressor, combustor, turbine, and
nozzle, and in addition monitors all the auxiliary systems such as air bypass, fuel
injector feed, structural vibrations, etc. Regardless, a robust and successful ACS
for combustion will go a long way towards achieving this objective, and this
chapter focuses on this particular area.
It is implicit in the discussion here that turbulent combustion is inherently an
unsteady, 3-D process involving nonlinear coupling of the acoustics of the system
with unsteady heat release and turbulent shear flow in the combustion chamber
(Menon, 2005; Poinsot et al., 1987; Zinn and Lieuwen, 2005; and further discus-
sion below). Thus, both CI and LBO are unsteady phenomena as well, and so any
successful ACS to control them will have to be a time-resolved system (Dowling
and Morgans, 2005). In fact, response in “real” time is a key requirement of any
ACS, and especially so for ACS used for combustion control, since the time-scale
of flame response to perturbations can be very small. Besides, failure to control CI
or LBO within milliseconds can be the difference between a commercially viable
and profitable stable system and catastrophic failure of an innovative design
(Lieuwen and McManus, 2002; Zinn and Liewuen, 2005). In addition to the need
for temporal fidelity in ACS implementation and response, spatial resolution is
also very important since sensors (and actuators) are physically located at specific
locations in the combustor and, therefore, the entire control strategy requires
merging spatially resolved sensing and actuation with a temporally resolved
response (Annaswamy and Ghoniem, 2002; Dowling and Morgans, 2005). Clearly,
these issues are valid for both experimental and numerical studies.
It is not possible nor is it intended to cover all these aspects and all the observa-
tions and discoveries by many teams and individual researchers. The references
cited in this chapter serve to acknowledge all these efforts (but are in no way
comprehensive), and only a few representative results are discussed in detail, for
324 S. MENON AND B. T. ZINN

brevity. Furthermore, although CI and LBO are the focus of this chapter, other
critically relevant physics are not discussed here but their importance is noted. For
example, pattern factor is a measure of the mean (and/or fluctuation intensity)
inflow conditions (e.g., temperature, species) from the combustor exit to the
turbine inlet, and is considered a fundamental design parameter since the perfor-
mance of the turbine is directly linked to the combustor exit conditions (e.g., Chen
et al., 1999; DeLaat et al., 2000; Palaghita and Seitzman, 2004, 2005; Tuncer
et al., 2005). Control of pollutants, primarily carbon monoxide (CO), oxides of
nitrogen (NO and NO2, together denoted NOx), unburned hydrocarbons (UHC),
and soot is also an important measure of the efficiency of the combustor design
(e.g., DeLaat et al., 2000; Mongia et al., 2001, 2003). Interestingly, control of CI
and LBO can have some influence on both of these design-critical physics.
This chapter is organized as follows. The next section will give an overview of
CI and LBO, followed by sections that will address active control of CI and LBO.
Both experimental and numerical studies will be addressed in these sections.
The chapter will conclude with a discussion of future research and development
needs in this area.

II. Physics of Combustion Instability and Lean Blowout


The physical processes behind CI and LBO are tied to the nonlinear interactions
between the various modes of wave motion in an unsteady compressible flow. As has
been pointed out (Chu and Kovasznay, 1958), in such a flow, acoustic wave motion
can interact with the unsteady heat release (and the associated volumetric expansion
effect), as well as with shear flow (vorticity modes and hence, turbulence) and mov-
ing “hot” spots (entropy modes and hence, temperature fluctuations) in a complex
manner. Some interactions are direct and stronger than others (e.g., acoustic wave
can interact with unsteady heat release even in the absence of turbulence). However,
in most power or propulsion systems the acoustic-vortex–flame (AVF) interactions
involve all these processes (see Lieuwen and Yang, 2005; Menon, 2005; Poinsot and
Veynante, 2005). It is apparent that if such interactions become coupled at character-
istic spatial and/or temporal time scales then situations can arise that may or may not
be stable. Both CI and LBO are manifestations of these AVF interactions.

A. Combustion Instability
Generally, combustion instability describes a situation in which the combustion
process excites oscillations of one of the natural acoustic modes of the combustor.
These acoustic oscillations are accompanied by periodic variation of the tempera-
ture, velocity and pressure fields within the combustor, resulting in a periodic
combustion process and thus a periodic heat release process. Combustion instabil-
ity has been major problem for many combustion devices for a long time (see
historical discussion in Culick and Yang, 1995; Zinn and Lieuwen, 2005). Although
the basic mechanism has been understood, albeit under simplified conditions, the
nonlinear complexity of this process in full-scale devices has continued to vex
practical devices. Rayleigh (1945) demonstrated that if unsteady heat release is in
phase with the acoustic (pressure) mode in the combustor then it will eventually
result in large-amplitude pressure oscillation that can either blowout the flame or
COMBUSTION CONTROL 325

cause structural damage due to excessive mechanical loads, vibrations, and/or


thermal stresses. Mathematically, this criterion can be written locally as:

Ú Ú p¢( x, t ) q¢( x, t ) dt dV ≥ Ú Ú [ L ( x, t ) + L ( x, t )] dt dV
Vc T Vc T
a d (1)

Here, p, q¢, La and Ld are respectively the unsteady pressure fluctuation, unsteady
heat release fluctuation, acoustic energy loss from the combustor volume Vc by
transmission or radiation through the inflow/outflow boundaries, and acoustic
energy loss due to viscous dissipation and other dissipative losses. The time inte-
gration in Eq. (1) is over a representative time period T of oscillation and the space
integration is over the entire volume of interest. The criterion suggests that if the
pressure and heat release fluctuations are in-phase and if the in-phase addition of
energy to the pressure fluctuations exceeds all the acoustic energy losses from the
system, then the instability will be driven. The equality in the above equation
indicates a global balance between energy addition and removal and, in this case,
a limit cycle of pressure oscillation is achieved (Lieuwen, 2002; Poinsot and
Veynante, 2005). Note that the final peak-to-peak amplitude of pressure oscilla-
tion can be substantial but even a 2 psia amplitude is sufficient to cause severe
structural damage (Lieuwen and McManus, 2002).
There are many variants of the above Rayleigh criterion. For example, if we
define a local Rayleigh parameter: R(x, t) = p¢(x, t)q¢(x, t) - [La + Ld](x, t), then the
preceding criterion is simply

Ú Ú R ( x , t ) dt dV ≥ 0
Vc T

In practical systems, it is quite difficult to fully estimate the losses locally and
only global criteria can be estimated. However, it is important to note that local
damping and/or driving can occur in both space and in time. Thus, various local
criteria can be defined (see for example, in Poinsot and Veynante, 2005; Chap. 8),
for example, a time-averaged criterion that varies in space:

R( x ) = Ú R( x, t ) dt
T

or a global volume-averaged criterion that varies in time:

R (t ) = Ú R( x, t ) dV
Vc

These parameters are useful in the interpretation of numerical simulation results


(e.g., Menon, 1992; Menon and Jou, 1991).
The instability driving mechanism can have many sources in the system design
and its operational envelope (Cohen and Banaszuk, 2005; Ducruix et al., 2005). As
noted earlier, all the three elements of AVF interactions can provide a mechanism for
driving combustion instability (CI). Past studies have identified many mechanisms,
326 S. MENON AND B. T. ZINN

such as: a) acoustic oscillation in the combustor exciting pressure oscillations in the
fuel feedline resulting in heat release fluctuations (Larson et al., 1981; Pandalai and
Mongia, 1998); b) imperfect mixing of fuel and oxidizer (in premixed system)
upstream of the combustor leading to time-dependent changes in the equivalence
ratio in the inlet (spatial variation in the inflow can also result from this imperfect
mixing) (Lieuwen et al., 2001; Richards et al., 1999); c) large-scale vortex shed-
ding from the flame holder and interactions of swirling shear layers with the flame
structure (e.g., Candel, 2002; Menon and Jou, 1991; Poinsot et al., 1987; Schadow
and Gutmark, 1992); and d) unsteady flame area changes that can contribute to a
periodic variation in heat addition (e.g., Candel, 2002). Other mechanisms are also
possible, such as droplet atomization and vaporization, interactions between multiple
fuel injectors, etc. (Coker et al., 2006; Lal et al., 2003b).
Thus, the Rayleigh’s criterion (Rayleigh, 1945) can be generalized to form
other similar criteria that show the actual physical processes involved (Poinsot
and Veynante, 2005, Chap. 8). For example, in the presence of mass sources, body
forces, or heat sources, all the following integral inequalities indicate driving of
combustion process or instability (assuming all losses are ignored): 1) acoustic
oscillations (p¢) driven by mass source oscillation (m·¢):

Ú Ú m ¢p ¢ / r
Vc T
0 dt dV > 0

2) acoustic oscillations (denoted by velocity fluctuation, v¢) driven by unsteady


(body or surface) forces in the system

Ú Ú F¢ ◊ v¢ dt dV > 0
Vc T

and 3) temperature fluctuations (T¢) driven by unsteady heat release (or heat
sources) (Q¢):

Ú Ú Q ¢T ¢ / T
Vc T
0 dt dV > 0

It should also be noted that the last integral is essentially equivalent to the original
Rayleigh’s criterion (Rayleigh, 1945), although here it uses temperature perturba-
tion while the “classic” Rayleigh’s criterion uses pressure perturbation. If the
above inequalities are “reversed” by some mechanism(s), then these interactions
have the potential to damp CI. This realization has played an important role in the
development of passive and/or active control approaches that, in part, attempt to
reduce the magnitude of the integrals or reverse their signs. Interestingly, success-
ful strategies used to keep campfires burning effectively in high wind (e.g., by
using wind breakers) or with weak burning fuels (e.g., by explicitly adding fast
burning secondary fuel) are perhaps early variants of passive or active strategies
that were developed using empirical observations.
To schematically describe the operation of various passive and active control
approaches, it is useful to examine the conditions under which combustion
COMBUSTION CONTROL 327

Fig. 1 Qualitative description of conditions under which combustion instability is


excited.

instability is spontaneously excited. Figure 1 indicates that acoustic oscillations


are driven in a combustor by a feedback-like interaction between the oscillatory
combustion process and oscillations of one or more natural acoustic modes of the
combustor. The effective driving of acoustic modes by the combustion process
requires that the characteristic time for the combustion process be of the same
order of magnitude as the acoustic time, which generally equals the period of the
most unstable (i.e., largest amplitude) mode. Note that, although in many of the
reported studies the most unstable mode is the longitudinal mode in the combustor
(hence, in the 100-Hz range), azimuthal and even helical modes (that are in the
kHz range) can be the most unstable, depending upon the device geometry
(Annaswamy and Ghoniem, 2002; Paschereit and Gutmark, 2002; Zinn and
Lieuwen, 2005). Furthermore, the heat release oscillations must occurs at com-
bustor regions where the amplitude of the acoustic pressure oscillations is large
and the magnitude of the phase difference between the heat release and pressure
oscillation must be smaller than 90 deg. The acoustic modes are also damped by
radiation and convection of acoustic energy out of the system (e.g., through the
nozzle), and by dissipation due to viscosity and heat conduction (Poinsot and
Veynante, 2005). An acoustic mode is spontaneously excited only if the energy it
receives from the combustion process exceeds the acoustic energy it loses due to
the various damping processes. The instability reaches its limit cycle (i.e., when
the amplitude of the instability no longer changes) when the acoustic energy sup-
plied and removed from the oscillation during a cycle equal one another. The
amplitude of the limit cycle oscillation, which often determines whether the insta-
bility can be tolerated or not, is controlled by nonlinear flow and combustion
processes involved in the AVF interactions.
328 S. MENON AND B. T. ZINN

Fig. 2 Growth of pressure oscillation leading to limit cylce behavior during combus-
tion instability (Poinsot and Veynante, 2005).

Figure 2 shows a typical limit cycle signature of pressure fluctuation (Poinsot


et al., 1988) during combustion instability. As the figure shows, the instability ini-
tially grows slowly with a linear growth rate; depending upon the system this
growth may rapidly transition into a highly nonlinear, exponential growth with the
resulting increase in the pressure oscillation. During the growth system failure is
possible (Lieuwen and McManus, 2002); however, the oscillation usually shows a
limit cycle behavior with a fixed amplitude. During this limit cycle oscillation the
energy added to the pressure oscillation is balanced by the acoustic losses from the
flow through the boundaries via viscous dissipation.

B. Lean Blowout
LBO is a flame blowout process that occurs when the flame is very lean.
Obviously, this process can be very detrimental to the operational status of the
engine. LBO implies total or global flame extinction; again, as in CI, even the
beginning of this process is unacceptable. Flame blowout can occur for many
reasons (especially in military engines undergoing violent maneuver). Safety con-
siderations dictate that combustors operate at a certain global equivalence ratio that
is sufficiently removed from that at which LBO occurs. However, since it is not
possible to maintain the same safety margin over the whole engine’s operating
envelope, current combustors must operate at “non-optimal” equivalence ratios at
some operating conditions, resulting in downgraded engine performance.
LBO can be simulated in almost any combustor setup in the laboratory by
simply reducing the fuel flow rate (Gutmark et al., 1991; Nair and Lieuwen, 2005,
2007). LBO can also occur when lean combustion systems are subject to sudden
and violent disturbance, for example, by a burst of turbulences or growth of
pressure disturbance (akin to the CI). Another interesting observation, at least in
premixed systems, is that, as the fuel equivalence ratio is reduced and the system
approaches its LBO limit, rapid increase in CO emission occurs (Bhargava et al.,
COMBUSTION CONTROL 329

2000). In some systems, this increase is also accompanied by increase in pressure


fluctuations and thus, depending on the hardware system, LBO and CI can occur
together. However, this coupling (which is related to fluctuations in unsteady head
release driving pressure oscillations) may not occur in all systems. Regardless,
this effect is related to the flame stability since in the near lean limit local flame
extinction can lead to an increase in unburned hydrocarbons (UHC) and this in
turn contributes to the increase in CO levels (Colby et al., 2006).
To highlight the emission characteristics near LBO, Fig. 3 shows a typical labo-
ratory scale combustor used to study physics in very lean premixed systems
(Bhargava et al., 2000). Figures 4a and b show respectively the CO and NO emis-
sion signature in this combustor (at a specified location) as a function of equiva-
lence ratio. Clearly, the reduced NO emission with decrease in equivalence ratio
is understandable since the flame temperature decreases, as the mixture is made
leaner. The CO level initially decreases but after reaching a minimum, it rapidly
increases as the equivalence ratio keeps decreasing. This rapid increase in CO for

Fig. 3 The DOE-HAT Combustor (Bhargava et al., 2000) used to study lean pre-
mixed combustion and emission characteristics. Notice that the emission sample
probe is located at only one location in the combustor.
330 S. MENON AND B. T. ZINN
LIVE GRAPH
LIVE GRAPH Click here to view
Click here to view 40

NOx at 15% excess O2, ppmv


40
CO at 15% excess O2, ppmv

100 psi 280 psi


200 psi 30
30 280 psi
20 200 psi
20

100 psi
10 10

0 0
0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.4 0.45 0.5 0.55 0.6 0.65
Equivalence ratio Equivalence ratio

Fig. 4 CO and NOx emission at the probe location in the DOE-HAT combustor for
different L operating pressures (Bhargava et al., 2000). In all cases, NOx emission
decreases with decrease in equivalence ratio but CO emission first decreases and then
rapidly increases very close to the lean flammability limit.

a very small reduction in equivalence ratio is characteristic of the beginning of the


LBO process. It is obvious that, for a given system, operating at the minimum CO
(which also has minimum NO) emission operating point would be ideal from an
emissions point of view. However, in reality, this minimum point is not known a
priori and is a function of many factors of the actual design. Even if this minimum
point is known, operating the combustor at this point is not attempted since even
a small excursion in the equivalence ratio can push the system towards the LBO
limit. (and rapidly increase the CO emission before reaching this limit).

III. Control of Combustion Instability


Most real engine systems are very complex, involving many parts. However,
very little effort has been reported on active control of full engine systems while
operating under realistic conditions. Real systems may also involve many injec-
tors, fuel and air feed systems that have to work together. Mimicking these features
in a laboratory scaled device is very difficult, not to mention expensive. Thus,
most experimental studies in the laboratory have been limited to single injector
systems at atmospheric conditions, although more recent studies are slowly
moving into high-pressure operating conditions.

A. Passive Control of Combustion Instability


Most passive control approaches modify the combustion process to reduce its
sensitivity to variation in the system performance envelope (e.g., take-off and
landing, and cruise) rather than an optimization within any particular operating
state. Typical passive control techniques involve changing the various subsystem
geometry, e.g., changing the injector design, changing the combustor geometry
by the addition of baffles, modifying the wall boundary conditions by adding
acoustic liners, etc. In most of these geometrical design changes, the primary
goal is to modify the instability mode’s frequency and/or amplitude so that it is
no longer in-phase with the system acoustic modes. Acoustic liners are used
to increase the damping process so that the energy addition to the pressure
COMBUSTION CONTROL 331

perturbation is reduced. In many applications all these changes are combined to


achieve operational control.
It can be noted that another school of thought is still focusing on developing
robust systems that do not require ACS to meet the design objectives. Earlier
successful passive strategies developed for liquid rocket motor combustion insta-
bility control (e.g., Cox and Marble, 1953; Crocco et al., 1960; Culick and Yang,
1995; Harrje and Reardon, 1972; Yang et al., 1990) fall in this arena. In fact, pas-
sive control strategy (e.g., Becker and Hassa, 2003; Bellucci et al., 2004; Dowling
and Dupere, 2005; Richards and Straub, 2005; Schadow and Gutmark, 1992;
Steele et al., 2000) has been and still remains a viable method to control CI even
in gas turbine engines because it is considered more reliable for field deployment.
Recent effort has also focused on developing and demonstrating alternate com-
bustion systems that can achieve low emission over a wide range of fuel types
without requiring an ACS. For example, researchers (e.g., Straub et al., 2005)
have been exploring combustion systems that can meet these goals without requir-
ing major hardwire redesign. Concepts that combine rich-quench-lean-burn
(RQL) with trapped vortex combustor (TVC) have demonstrated the potential for
high-powered systems capable of delivering the design output without high emis-
sion and/or instability. Another system that can also reduce emission without
exhibiting any major instability or LBO is the lean direct injection (LDI) combus-
tion system proposed by NASA (e.g., Tacina et al., 2003). Studies (Becker and
Hasse, 2003; Lal et al., 2003b) have demonstrated that with proper droplet size
distribution and flow pattern it is possible to avoid combustion instability and also
to modify the emission pattern.
Fine-tuning and fully optimizing such systems remains an elusive goal at pres-
ent. Not only do we need detailed measurements over wide operating conditions
but also a comprehensive predictive capability that can help optimize such designs.
Predictive capability is critically needed because extending laboratory scale
devices (that typically operate under atmospheric conditions) to high-pressure
production systems is a major issue; new physics can manifest itself at high pres-
sure and detailed non-intrusive measurements are very difficult in actual opera-
tional combustors.

B. Active Control of Combustion Instability


Within the active control strategy, two approaches, namely open-loop and
closed-loop control are used for fundamental and practical demonstrations. It is
not possible to cover all the features and flavors of these approaches; therefore this
chapter focuses on only a subset of these studies. However, an attempt is made to
provide many key references to past and current studies. To discuss ACS for a
combustor, consider the generic closed-loop system shown in Fig. 5. It consists of
sensor(s), an observer (or an analyzer), a controller and an actuator(s). An open-
loop system may at minimum consist of a controller and actuator(s). Figure 6
shows an example of the setup for a control system used in a dump combustor that
includes both open- and closed-loop systems.
From a practical point of view, open-loop control actuation can only be imple-
mented at specific locations (this is also true for closed-loop control) but their
impact on the entire flowfield is critical to the understanding of its relevance.
332 S. MENON AND B. T. ZINN

Fig. 5 Schematic of a typical active control system used for combustion control.

Such studies serve at least three important purposes: a) they can be used to evalu-
ate the response of the system to “forced” instabilities at various frequencies, and
hence can be used to map out the system response (in some cases such maps are
used to verify and/or demonstrate an ACS system); b) when excited at the relevant
resonant frequencies these studies can be used to determine if CI and/or LBO can
be excited in the first place and c) such studies, when properly documented and
characterized, can be a boon for simulation model validation.

1. Experimental Studies of Control of Combustion Instability


There have been many open-loop studies reported in literature. In general, these
studies provide great insight into the response of the combustion system to specific
inputs into the AVF interaction process. Both open- and closed-loop systems
employ similar control and actuation strategies, and attempt control of one or
more of the physical processes in the AVF interactions. Many of these studies are
discussed in detail in Dowling and Morgans (2005). For example, the following
have all been used in both open- and closed-loop control systems: shear layer
excitation that impacts fuel–air mixing and shear layer vortex roll-up/pairing/
breakup (e.g., McManus et al., 1990; Schadow and Gutmark, 1992); fuel and/or

Fig. 6 Schematic of an ACS with both open- and closed-loop setup in a dump
combustor.
COMBUSTION CONTROL 333

air flow rate modulation that impact fuel–air mixedness, equivalence ratio, and
flame structure (e.g., Cohen et al., 2001; Neumeier et al., 1997; Richards et al.,
1999, 2007; Uhm and Acharya, 2005); acoustic forcing using loud speakers that
impacts pressure fluctuations (e.g., Bloxsidge et al., 1988; Paschereit and Gutmark,
2002); spray property changes in liquid-fueled systems that impact droplet disper-
sion, vaporization, and gaseous fuel–air mixing (e.g., Becker and Hassa, 2003;
Hermann et al., 1996; Lal et al., 2003b; Yu and Wilson, 2002; ).
Examples of actuators used to achieve these physical property changes (note
that these same actuators can be and are used in the closed-loop studies as well)
include valves that oscillate the air (or fuel) flow rate into the combustor (e.g.,
Bloxsidge et al., 1988; Tuncer et al., 2005; Uhm and Acharya, 2004), speakers
that excite acoustic oscillations with desired phase and amplitude within the com-
bustor (e.g., Paschereit et al., 1999), synthetic jet actuators (Ritchie et al., 2000)
that introduce vortical features at the small scale to enhance mixing, and second-
ary or primary fuel injection manipulations (e.g., Auer et al., 2005; Barbosa et al.,
2007; Cohen and Rey, 1999; Cohen et al., 2001; Ghoniem et al., 2005; Gutmark
et al., 1998; Hathout et al., 2002; Jones et al., 1999; Kim et al., 2000; Lal et al.,
2003a, 2004; Lang et al., 1987; Langhorne et al., 1990; Richards et al., 1999; Yi
and Gutmark, 2007a, b; Yu and Wilson, 2002; Yu et al., 1996; Zinn and Neumeier,
1997) that modulate the injection rate of all or a fraction of the supplied fuel. As
this reference list suggests, controlling the fuel flow rate or modulating the fuel
(primary or secondary) has become a major area of ACS for combustion control
since this approach directly targets the source of combustion. Furthermore, since
the amount of fuel is substantially much less than the air flow, and fuel injection
locations are generally fixed by operating requirements, fuel control is easier to
implement in practical systems. These features have been exploited in all types of
combustion systems (premixed, non-premixed, and spray). Cyclic and/or con-
trolled fuel injection at specific locations. In-phase or out-of-phase with the local
pressure fluctuation, they have been demonstrated to be a viable active control
strategy in all these systems.
Open-loop control approaches involving open-loop pulsing at a frequency that
differs from the instability frequency, combustion time control, hysteresis control,
and actuation at a sub-harmonic frequency of the instability have all been demon-
strated in recent years. Fuel injection rate pulsed at a frequency that differs from
the instability frequency (e.g., Lubarsky et al., 2004; Prasanth et al., 2002; Richards
et al., 1999, 2003) and it has been shown that this approach can significantly attenu-
ate the instability when the forcing occurs at selected frequencies. However, these
frequencies must be determined in advance, and this is potentially a drawback. The
“combustion time control” approach (Conrad et al., 2004) modifies the character-
istic droplet evaporation time to the order of the acoustic period of the instability.
Conrad and co-workers employed a “smart” liquid fuel injector with the capability
to modify the combustion time by changing the spray characteristics. This study
shows that a “one-time action” that modifies the fuel spray characteristics can sig-
nificantly reduce the amplitude of the instability. The one-time need and subsequent
stable operation suggests that if the optimal controller (or “smart” fuel injector)
setting can be determined in advance, controller actions, which modify the com-
bustion process can be implemented “one at a time” to prevent combustion insta-
bility whenever the engine’s operating conditions change.
334 S. MENON AND B. T. ZINN

An approach for controlling combustion instabilities based upon hysteresis


control has also been proposed (Knoop et al., 1977). It exploits the fact that
unstable combustors may exhibit hysteresis, involving operations with high- and
low-amplitude oscillations at the same equivalence ratio. The control approach
involves subjecting the combustor to external disturbances that “move” its opera-
tion from the large-amplitude to the low-amplitude branch of hysteresis. Active
control of combustion instabilities by forced variations of the equivalence ratio
has been reported (Richards et al., 1999) in combustors that exhibit unstable oscil-
lations over a limited range of equivalence ratios.
In the closed-loop strategy (Figs. 5 and 6), the two additional elements are the
sensor(s) and the analyzer systems. The sensors may include one or more trans-
ducers that measure, e.g., dynamic pressure and/or flame chemiluminescence
(which can be correlated with the time dependence of the reaction rate). The sig-
nals measured by these sensors have frequency spectra that describe the character-
istics of the flame response to its environment (e.g., turbulence, acoustic waves,
inflow, and boundary conditions). The measured data are then analyzed by an
“observer” (Neumeier and Zinn, 1996a, 1996b, 1996c), or by a model (either
preset or dynamic) of the problem (Annaswamy and Ghoniem, 2002; Annaswamy
et al., 1998; Hathout et al., 2002; Morgans and Dowling, 2007; Morgans and
Stow, 2007; Murugappan et al., 2003; Riley et al., 2004), to determine the “state”
of the system in terms of (for example) the frequencies, amplitudes and phases of
the various system responses (e.g., excited acoustic modes, turbulence enhance-
ment or suppression). These data are then used by the controller, which may use
one of several approaches (e.g., model-based, adaptive, neural net, fuzzy logic, or
rule-based) to determine the control signal(s) for the actuator(s) whose task is to
modify the combustor operation in a manner that would eliminate the problem,
i.e., damp the instability or prevent LBO.
A brief (not comprehensive) review of highlights from some of these past stud-
ies is now given. Padmanabhan et al. (1995) developed a strategy to optimize the
performance of laboratory scale combustion to maximize volumetric heat release
and minimize pressure fluctuations. The control system simultaneously sensed
and controlled both these variables by combining spanwise forcing of the inlet
boundary layer with crossflow jets upstream of the inlet. Results showed the
control strategy could continuously seek an optimal performance even when the
inlet condition changes (such as flow disturbances) are not known. Uhm and
Acharya (2004, 2005) have studied techniques for controlling CI using a high-
momentum air jet that is used to change the mixing process in a swirl-stabilized
spray combustor. They showed that if the jet is injected in the regions of positive
Rayleigh index (the local source of instability initiation) then control can be
maintained.
Acoustic feedback and secondary fuel injection control have been two major
(and successful) approaches for control of CI. However, the strategies and the
nature of injection vary from group to group, and also from test case to case.
Earlier studies focused on gas fueled (either in premixed or non-premixed state)
systems to model dump combustors that were considered representative of ramjets
and/or afterburners (e.g., Billoud et al., 1992; Candel, 1992; Candel and Poinsot,
1987; Poinsot et al., 1987; Schadow and Gutmark, 1992; Yu et al., 1996). Studies
showed that controlled pulses of secondary fuel injection can be used to achieve
COMBUSTION CONTROL 335

good control authority. These observations have been verified and re-demonstrated
in more current systems.
Many control studies of liquid fueled systems under lean conditions have been
reported in dump combustors (e.g., Cohen and Rey, 1999; Hermann et al., 1996;
Yu et al., 1996), and in more complex multi-swirl systems (e.g., Gutmark et al.,
1990, 1998; Johnson et al., 2001a, 2001b; Lal et al., 2004; Nabi et al., 2000; Yi and
Gutmark, 2007a, 2007b). Both open- and closed-loop studies have been reported,
and in nearly all of these studies a model-based controller has been employed.
Other approaches that employ empirical and knowledge-based models have also
been explored, although in a limited sense. Allen et al. (1993) demonstrated a
closed-loop neural network (NN)-based control system using time-resolved imag-
ing of chemically specific emission patterns in the flame. They showed that purely
spatial images could be successfully interpreted by NN to achieve control (DeLaat
et al., 2000; Gutmark et al., 1990). Fuzzy logic-based active control of combustion
instability has also been demonstrated in laboratory scale devices (e.g., Coker
et al., 2006; DeLaat et al., 2000; Nelson and Lakany, 2007). An interesting obser-
vation in liquid fueled systems is that the fuel droplet size distribution (and perhaps
its injection process) can be used to control combustion dynamics. From a practical
point of view, spray injection manipulation in swirling air provides a means to
control both fuel–air mixing and flame stabilization processes.
As a demonstration example, a particular ACS system for control of CI is now
discussed to highlight the application. Here, we consider an ACS for applications
in a combustor that burn liquid or gaseous fuels, and require nearly instantaneous
attenuation of large amplitude combustion instability whose characteristics are not
known a priori and may vary in time (Neumeier and Zinn, 1996a, 1996b, 1996c).
One successful configuration is shown in Fig. 7, and will be referred to as combus-
tion instability active control system (CIACS) in this discussion. In this test setup,
ACS is installed in a combustor that burns natural gas in air at a mean pressure
of ~6.8 atm. Eighty percent of the fuel is premixed with all the air and injected into
the combustor through a number of orifices uniformly distributed around the
periphery of the injector plate with their axis inclined relative to the combustor
axis. Periodic modulation of 20% of the fuel injection rate is used to control the
instability. This fraction of the fuel is injected through a second set of orifices that
direct the fuel jets radially towards the location of the combustion zone.
The only sensor used in this CIACS is a pressure transducer installed at the
upstream end of the combustor. The choice of this location is based on the
observation that the maxima of all the combustor’s axial acoustic modes occur
at that location. An observer that operates in real time uses the pressure data and
determines the characteristics of a specified number of the “most unstable”
combustor modes. The controller uses the observer’s information to define the
control signal (e.g., gain, phase-shift) for each of the observed modes. The
control signal information is obtained from a stored database on the open-loop
flame response data. The generated, time-dependent control signal (for each
mode) are synthesized into a single control signal that is sent to a fuel injector
actuator that then modulates the injection rate of the fuel jets. This control
approach employs the Rayleigh’s criterion so that the phase of the control signal
of each mode is 180 deg out of phase with respect to the pressure oscillations.
The CIACS continuously monitors the conditions within the combustor and can
336 S. MENON AND B. T. ZINN

Fig. 7 Schematic of an ACS developed and demonstrated at Georgia Tech for


combustion instability control (Neumeier and Zinn, 1996b).

change the gains and phases of the control signals provided by the controller if
the characteristics of the instability change in time.
The novel feature of this CIACS is its observer. This observer is a mathematical
algorithm (not a model of the system) that analyzes the measured pressure (or any
other) signal by use of a wavelet-type transform (Farge, 1992) with feedback to
identify the amplitudes, frequencies, and phases of a pre-specified number of the
most unstable combustor modes in real time. Essentially, the observer first identi-
fies the mode with the largest amplitude, followed by the identification of the
mode with next largest amplitude and so on. This is accomplished without any
prior knowledge of the natural acoustic models of the combustor. The fuel injec-
tor actuator is a magnetostrictive actuator that can modulate the injection rate of
a gaseous or a liquid fuel stream over a 0–1500 Hz frequency range with large
flow rate.
An example of the closed-loop performance of the CIACS, when it controls
only the most unstable combustor mode, is shown in Fig. 8. It shows the time
dependence and FFT of the combustor pressure before and after activation of the
control system. Figure 8a shows that large amplitude (i.e., 15 psi, peak to peak,
which approximately equals 15% of the mean combustor pressure) limit cycle
oscillations are nearly completely damped in 40 ms. Furthermore, comparison of
the FFTs of the combustor pressure oscillations before and after control activation
(Figs. 8b and 8c) shows that the amplitude of the fundamental acoustic mode is
reduced by 26 dB, and that once the fundamental mode is damped, all its harmon-
ics are also damped. These results show that a properly designed ACS could
effectively damp large amplitude instabilities nearly instantaneously and thus
avoid damaging the engine or stopping operation.
COMBUSTION CONTROL 337
LIVE GRAPH
Click here to view

LIVE GRAPH LIVE GRAPH


Click here to view Click here to view

Fig. 8 Performance of the GTAGS in the combustor setup shown in Fig. 7. a)


Combustor pressure signal; b) FFT of the pressure signal without control; and c) FFT
of the pressure signal with control (Neumeier and Zinn, 1996b).

Obvious drawbacks of this CIACS are that the pressure sensor is located at an
“optimal” location and that its controller requires data that must be obtained in
separate open-loop tests, which can be expensive and time consuming. The latter
problem can be overcome by use of adaptive controllers which have been demon-
strated in the literature (e.g., Billoud et al., 1992; Docquier and Candel, 2002;
Johnson et al., 2001a; Padmanabhan et al., 1995). These adaptive controllers
determine the necessary control signal in the course of applying active control
and offer considerable potential, although their use increases the time required to
actively control the instability.
Studies with a large-scale sector combustor under realistic operating conditions
(e.g., Gutmark et al., 1998; Johnson et al., 2001a, b; Yi and Gutmark, 2007a, b),
and full-scale engines (e.g., Hermann and Hoffman, 2005), have demonstrated
that ACS can damp instabilities in practical systems. Hermann and Hoffman
(2005) describe the combined use of passive and active means to control a multi-
mode tangential instability in a full scale, 260-MW gas turbine. In this case, an
ACS consisting of a pressure sensor, a controller, and a valve is installed in each
of the 24 burners and used to damp the instability by modulating the injection rate
of the pilot fuel. This system operated successfully for (18,000 h). Such demon-
stration provides some confidence that pressure transducers and software devel-
oped in laboratory studies could be used, perhaps with some minor modifications,
338 S. MENON AND B. T. ZINN

in full-scale engines, and that available fuel injector actuators could be scaled up
to modulate the large fuel flow rates (into the pilot or main combustion regions)
needed for active control of combustion instabilities in practical systems.

2. Numerical Studies of Control of Combustion Instability


Most of the earlier numerical efforts were focused on simple laboratory scaled
rectangular or axisymmetric dump combustors. Due to the inherent unsteady
nature of the processes, time-accurate simulations are needed, and in the early
effort, 2-D or axisymmetric approximations are necessary due to computational
resource limitations. Nevertheless, these studies attempted large-eddy simulations
(LES) using the flame sheet (or the G-equation) approach with a subgrid flame
speed model (e.g., Menon, 1992, 1995; Menon and Jou, 1991; Poinsot et al.,
1988), and demonstrating not only that CI can be simulated but also that active
control of the CI can be achieved using appropriate forcing techniques, such as
acoustic speakers (Menon, 1992) or secondary fuel injection (Menon, 1995).
More recent studies have focused on using full 3-D simulations (Brooks et al.,
2001; Selle et al., 2006; Stone and Menon, 2002, 2003). Model-based controllers
have been demonstrated in some simulations (e.g., Annaswamy and Ghoniem,
2002; Hathout et al., 2002; Menon and Yang, 1993). Kaufmann et al. (2002)
showed that to carry our proper modeling of combustion dynamics the inflow
must be properly simulated results. Simulation of combustion dynamics and its
control typically involve a computational domain that is only a subset of the actual
test facility or rig. Thus, the numerically imposed inflow and outflow conditions
can play fundamental roles in the actual physics being simulated.
Many studies in the past have also attempted to develop simplified models for
the combustion system and to develop algorithms that can be used in the simula-
tion. For example, 1-D models based on stability analysis (e.g., Dowling and
Stow, 2005; Yang and Culick, 1986; Yang et al., 1990) have been very popular for
the understanding of AVF interactions. Specifying appropriate boundary condi-
tions for such models can be problematic, and attempts have been made to use
transfer functions to relate acoustic perturbations in inlets, ducts, and fuel feed-
lines to the processes in the combustor (e.g., Bray et al., 2005; Hoffmann et al.,
2002; Lieuwen and Yang, 2005).
Hong et al. (2000) discuss a feedback controller to suppress CI in a dump
combustor using distributed actuators including modeling uncertainty. This is a
modeling/simulation study that employs an observer for robust estimation of
combustion dynamics and a H•-optimization technique for control. Extension
of this technique to a two-layer control system using multiple time-scale model
is reported in Hong et al. (2002) where secondary fuel injection is used
for control.
Shinjo et al. (2007) recently employed LES of secondary fuel injection control
in a lean premixed combustor that demonstrated large-amplitude pressure oscilla-
tion without control. It is shown that both constant and harmonic feedback injec-
tion reduced the pressure fluctuation level, with the latter method showing superior
ability. Selle et al. (2006) combined a LES solver with a Helmholtz solver to
investigate the nature of rotating modes in an industrial swirled burner. They
showed that, although the rotating motion in the chamber of the reacting case
COMBUSTION CONTROL 339

Fig. 9 Axial velocity contours and flame (thin white line) in a swirling premixed
dump combustor: a) low swirl and b) high swirl cases.

exhibits features similar to the classical hydrodynamic vortex breakdown process,


a careful analysis of the modes in the combustor showed that the source of this
rotating instability was primarily acoustics.
Other studies employed full 3-D LES studies to investigate open-loop control
of CI using either swirl or fuel modulation in premixed fueled dump combustors
operating under realistic conditions of a GE LM6000 type combustor (Stone and
Menon, 2002, 2003). Figure 9 shows the axial velocity contours in the near field
of the dump plane (only a 2-D slice of one-half domain is shown for clarity). The
low inlet swirl case (Fig. 9a) shows that the flow from the inlet is like a jet flow
and a large recirculation bubble forms at the base of the step. In this case, the
flame (shown as a thin white line) is attached at the step corner and looks more
like a jet flame. In the high swirl case (Fig. 9b) the swirl-induced vortex break-
down bubble (VBB) creates a large recirculation region around the centerline and
the flame is short and stabilized upstream of the bubble. The recirculation region
in the base of the step is much smaller and is pushed close to the step wall. As is
well known, swirl-stabilized flames are typical in all gas turbine combustors.
Thus, when instability occurs, both the VBB and the flame structure changes
rapidly, indicating a complicated coupling between fluid dynamics, flame struc-
ture, and the pressure pulsation (e.g., Menon, 2005).
Analysis of the above noted simulations show that the Rayleigh parameter is
aligned with the pressure signal in the low swirl case while it is more out-of-phase
for the high swirl case. These results are shown for a time window in Fig. 10. The
pressure trace also shows a higher peak-to-peak fluctuation in the low swirl case
indicating that relatively speaking this case is more susceptible to instability than
the high swirl case.
340 S. MENON AND B. T. ZINN LIVE GRAPH
Click here to view

Fig. 10 Comparison of the volume-averaged Rayleigh parameter and the pressure


fluctuation for the two swirl cases.

Open-loop studies were then conducted using both the high and low swirl cases.
In the first set of studies, the inflow swirl was increased gradually from the low
swirl to the high swirl conditions (Fig. 11a), while in the second case, the inflow
equivalence ratio was decreased from the high (stable) limit to a lean limit
(Fig. 11b). Both studies clearly show the impact of changing the inflow condi-
tions. With increase in swirl the instability (or peak-to-peak pressure fluctuation
level) decreases while decreasing the equivalence increases the pressure fluctua-
tion. As can be seen, swirl control takes nearly 15 cycles for the overall system to
respond, whereas the effect of equivalence ratio reduction is rapidly felt within
three cycles in the system. Although these studies did not actually control instabil-
ity, the fast system response to fuel mixture changes is consistent with the experi-
mental observations that fuel modulation or injection control is very effective in
controlling instabilities.
Finally, for a more visual demonstration of fuel modulation control, Fig. 12 shows
two instantaneous snapshots of the flame, vortex rings, and product temperature,
along with the pressure signal during control. In Fig. 12a the inflow fuel mixture
equivalence ratio is increased (rich mixture) and as a result, the flame temperature
increases, resulting in a hotter core region (red temperature contours). At this stage
of the control process, the pressure fluctuation begins to decrease (first quarter of the
p¢(t) signal). In Fig. 12b, the fuel mixture entering the combustor has become lean
and this results in a decrease in temperature (blue core region), and the pressure
oscillation increases. This time corresponds to the far right of the p¢(t) signal.
LIVE GRAPH COMBUSTION CONTROL 341
Click here to view

LIVE GRAPH
Click here to view

Fig. 11 Pressure fluctuation change in time with open-loop control: a) swirl control
and b) inflow equivalence ratio control (Stone and Menon, 2002, 2003).

IV. Control of Lean Blow Out


Lean premixed combustion is currently used in land-based gas turbines to burn
fuel at low temperature and thus reduce NOx emission. Lean prevaporized pre-
mixed (LPP) combustion has been advocated for aircraft engines to reduce NOx
emission as well. In both cases, the fuel is burned in a premixed combustion pro-
cess that occurs near the combustor’s LBO limit. Since the combustor is continu-
ously subjected to various disturbances, e.g., in air and fuel velocities, it cannot
operate safely too close to the LBO limit because a sufficiently large disturbance
342 S. MENON AND B. T. ZINN

Fig. 12 Instantaneous view of fuel modulation control. Shown are the flame surface
(black iso-surface), vortex rings (gray rings), product temperature (contours), and
the fluctuating pressure trace.

may cause the combustion process to cross the LBO limit, thus causing extinction
and stoppage of engine operation. Ideally, it would be desirable to operate the
engine with the same LBO safety margin at all operating conditions. Unfortunately,
existing technologies cannot provide such capabilities and this area remains a
major research focus at various laboratories.

A. Experimental Studies in Control of Lean Blowout


McManus et al. (1993) showed earlier that flow excitation can be used to control
premixed combustion in a simple 2-D dump combustor. Forcing by using frequen-
cies that corresponded to resonant and off-resonant vortex shedding frequencies
was studied and it was shown that the flame structure could be modulated with an
increase in CH emission up to 15%. In addition, the LBO equivalence ratio limit
was reduced by 6% and NOx emission were reduced by 20%. As in other nonre-
acting studies, the forcing frequency/amplitude and location are important param-
eters and also sensitive to the actual test facility geometry and conditions.
Here, we briefly discuss the effort to develop an ACS to prevent LBO at all
operating conditions (Muruganandan et al., 2005; Nair and Lieuwen, 2005, 2007;
Nair et al., 2004; Prakash et al., 2005; Thiruchengode et al., 2004; Zinn, 2005).
The characteristics of the developed ACS (denoted LBOACS) are described in
this section. The LBOACS shown in Fig. 13 is installed in a swirl-stabilized
combustor. Its fuel is supplied in two streams that are premixed with air upstream
COMBUSTION CONTROL 343

Fig. 13 An ACS used for LBO control in a swirl-stabilized dump combustor with
pilot fuel injection.

of the combustor. Most of the fuel is supplied into the main combustion region
through an annular opening near the injector periphery and a small fraction of the
fuel is supplied into a central pilot region. The LBOACS consists of acoustic and/
or optical sensors, a controller, and a fuel split actuator that controls the relative
fuel flow rates into the two combustion regions. The acoustic and optical sensors
measure respectively, the time dependence of the combustor dynamic pressure
and the flame chemiluminescence from a specific combustor region, respectively.
The controller analyzes the measured signals to detect specific features (i.e., LBO
precursors that describe local “events” consisting of flame extinction and reigni-
tion at a given combustor location) that are indicative of the onset of LBO and can
thus be used to determine the “closeness” of the combustion process to the LBO
limit. Once the controller determines that the combustion process is too close to
the LBO limit, it initiates or increases the fuel flow rate to the pilot region to pre-
vent LBO. Essentially, the controller controls the ratio of the fuel flow rates to the
main and pilot combustion regions.
Figure 14a shows the experimental setup that consists of a pressure transducer
and an optical fiber connected to a photomultiplier to measure the combustor
noise and flame chemiluminescence from a specific combustor region, respec-
tively. The measured data is continuously analyzed to detect for the possible
presence of LBO precursors. Additionally, it uses a thermocouple to measure the
quartz combustor wall temperature, and a sampling probe to collect a small flow
rate of products for chemical analysis. Figures l4b and c present two images of the
combustion zone taken at two different equivalent ratios. They clearly show that
the characteristics of the combustion process significantly change as the equiva-
lence ratio f decreases; i.e., combustion in the recirculation regions in the corners
at the bottom of the combustor disappears as f decreases from 0.85 to 0.79.
In this study, analyses of high-speed movies of the combustion region at differ-
ent values of f reveal that at low values of f local extinguishment and reignition
of the combustion process occur. It is also noted that the frequency of these
344 S. MENON AND B. T. ZINN

Fig. 14 Experimental system used for LBO control. a) Sensor and swirl flow assem-
bly; b) flame image at an equivalence ratio of 0.85; and c) flame image at an equiva-
lence ratio of 0.79 (Muruganandam et al., 2005).

extinguishment/reignition events (to be referred to as events in the remainder of


this discussion) increases as the equivalence ratio f decreases and the combustion
process approaches its LBO limit. Finally, it was discovered that these events can
be detected in measured sound and OH (and other radicals, e.g., CH) chemilumi-
nescence data by use of measured data filtering in combination with, for example,
wavelet, Fourier or statistical analyses.
To illustrate one of the developed data analysis approaches, consider the OH
radiation data presented in Fig. 15. It shows that the mean value of the measured
OH radiation, which is proportional to the mean reaction rate, decreases as the
equivalence ratio decreases. This figure also indicates the presence of “bursts” (or
spikes) in the OH data whose frequency increases as the equivalence ratio
decreases. Correlations of the measured OH radiation data with images obtained
with high-speed photography of the combustion process indicate that the observed
“bursts” are related to extinction/reignition events. The analyses also showed that
such an event occurr when the magnitude of the OH signal drops below one or two
threshold OH values that were arbitrarily chosen as some percentage of the mean
OH radiation value (Muruganandam et al., 2005; Nair et al., 2004). In this example,
an event starts and terminates when the magnitude of the OH radiation drops and
rises below and above two threshold lines that equal 0.65 and 0.5 times the mag-
nitude of the mean OH radiation, respectively. Since the occurrence of events is
related to measured mean values of OH radiation (or measured mean reaction
rate), their identification is independent of the engine’s power setting and possible
sensor drift due to aging and soot/dirt deposition. It can be also shown (Nair and
Lieuwen, 2005), that the optimal values of the chosen threshold values (relative to
the mean values) depend upon combustor design, sensor placement, and expected
LBOACS performance.
COMBUSTION CONTROL 345

LIVE GRAPH
Click here to view

Fig. 15 OH radiation signal for different equivalence ratio (Nair et al., 2004).

Figure 16 shows the typical dependence of the frequency of the events (deter-
mined from measured acoustic and OH radiation data) on the equivalence ratio, f.
It shows that the frequency of these events starts increasing at f ~ 0.76 as the LBO
limit at f ~ 0.75 is approached. The rate of this increase is initially small but rap-
idly increases as f further decreases. Figure 16 also shows that the frequency of
events provided by analyses of acoustic and OH radiation data are very close,
indicating that either one could, in principle, be used to detect the proximity of the
combustor to the LBO limit. When the LBOACS controller detects an increase in
the frequency of events similar to that shown in Fig. 16, it commands the actuator
to divert some of the fuel from the main combustion region to the pilot flame
region and thus stabilize the combustion process. A demonstration of such LBO
control is provided in Fig. 17. It describes a controlled experiment in which the
fuel flow rate was kept fixed and the airflow rate was suddenly increased to reduce
the equivalence ratio to below the nominal LBO limit of the combustor shown in
Fig. 13. Figure 17a shows that the controller detects an increase in the “alarm”
count (i.e., event frequency) shortly after the equivalence ratio starts decreasing,
while it is still above the LBO limit. As soon as alarms were detected, the control-
ler diverts some of the fuel into the pilot region to prevent LBO, as shown in
Fig. 17b. In fact, a fuel fraction between 14 and 18% of the total fuel input is
required to allow the combustor to operate below its nominal LBO limit without a
problem. The alarms disappeared and the LBO controller shut down as soon as the
equivalence ratio increased sufficiently above the combustor’s LBO limit. It is
noteworthy that these studies have also shown that since the LBOACS allows the
combustor to operate at globally lower equivalence ratios, and thus lower flame
temperatures, it could be used to lower the combustor’s NOx emissions.
To determine whether the developed optical sensing approaches could
be used in aircraft engines, the developed OH sensing approach was used to
346 S. MENON AND B. T. ZINN
LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view

Fig. 16 Use of threshold values to detect events. a) Percentage of mean OH chemi-


luminescence signal to define an extinguishment event and b) dependence of the fre-
quency of events on equivalence ratio (determined from the OH signal and acoustic
data measured in the setup shown in Fig. 14) (Zinn, 2005).

measure the onset of LBO in an atmospheric aircraft engine combustor sector


that used an aircraft engine liquid fuel injector and preheated air (Fig. 18). In
this study, the time dependence of the OH radiation from the combustion zone
was measured using an optical fiber and a photomultiplier as the fuel input rate
into the combustor was continuously decreased, eventually causing the com-
bustor to extinguish. In parallel, the time dependence of the total light radiation
from the whole combustion zone was measured by a high-speed camera and
COMBUSTION CONTROL 347

LIVE GRAPH
Click here to view

Fig. 17 Performance of the LBOACS in response to sudden decrease in the equiva-


lence ratio to below the nominal combustor LBO limits (Zinn, 2005).

compared with the measured OH radiation to determine whether the OH sens-


ing could be used to detect the proximity to the LBO limit in an aircraft com-
bustor, in spite of the fact that it only measured radiation from a fraction of the
combustion region.
The time dependence of the OH radiation measured during this study is shown
in Fig. 19 along with an instantaneous image of the flame. This signal describes
OH radiation from the encircled combustion region at the top of the photo, which

Fig. 18 Typical swirl cup single sector of commercial aircraft engine used to study
LBO in liquid fueled combustor in the laboratory (Zinn, 2005).
348 S. MENON AND B. T. ZINN

Fig. 19 Time-dependent behavior of OH signal measured by the optical sensor


(bottom) and an instantaneous image of the flame (Zinn, 2005).

shows an instantaneous view of the whole flame obtained by the high-speed cam-
era. The intensity of the measured OH radiation decreased as the fuel input into
the combustor decreased, eventually going to zero when the flame extinguished.
The OH signal became null exactly at the instant when the light intensity measured
by the high-speed camera became zero, indicating that the two optical measures
of the presence of the flame were consistent. The OH signal at the lower part of
Fig. 17 also includes events which are referred to as (LBO) “precursors” in the
figure. Analysis of this signal showed that the frequency of these events increased
as the LBO limit was approached. The results of this study indicate that the
developed optical sensing techniques could be potentially applied in aircraft
engines and land-based gas turbine combustors that burn liquid fuels to detect in
advance the onset of LBO.

B. Numerical Studies in Control of Lean Blowout


So far, no real demonstration of LBO and LBO control in numerical simulation
has been reported (at least to the authors’ knowledge). Unlike CI and its control,
COMBUSTION CONTROL 349

simulating LBO requires that all the features of the combustor and the combustion
process be properly captured. In particular, finite-rate kinetics has to be included so
that proper extinction/reignition process can be simulated, and also the flow–flame
coupling has to be correct to capture the physics near LBO. Finally, LBO is rather
sudden with rapid increase in CO emission as noted earlier, and thus, simulation very
close to the LBO limit must show these physical features without numerical
corruption.
There has been some progress towards modeling extinction/reignition processes
(Pitsch, 2006) but their actual application for LBO remains to be demonstrated.
Some recent efforts (Eggenspieler and Menon, 2004, 2005) have demonstrated
that with proper sub-grid modeling it is possible to capture the change in CO
emission as LBO approach is approached. Figure 20 shows CO and NO emission
prediction as a function of equivalence ratio in the DOE-HAT combustor discussed

LIVE GRAPH
Click here to view

LIVE GRAPH
Click here to view

Fig. 20 Prediction of CO and NO emission in the DOE-HAT using a sub-grid mixing


and combustion model (Eggenspieler and Menon, 2005).
350 S. MENON AND B. T. ZINN

earlier. To carry out these LES, a reduced reaction kinetics model and some other
simplifications are needed. Nevertheless, these studies show that it is possible to
study LBO using LES. However, active control of LBO using LES or any other
simulation approach still remains to be demonstrated.

V. Future Prospects
There are many unresolved issues related to ACS performance in practical sys-
tems and research is still underway in many laboratories. Clearly, the optimal per-
formance of an ACS strongly depends on the characteristics of its components, i.e.,
sensor(s), observer, controller and actuator(s), and their interactions. Using only a
single pressure transducer (as most ACS have done to date) may not always work
since without a priori open-loop studies, it could be “accidentally” installed at a
location where the amplitude of the excited pressure oscillations is minimum (or
zero). Practical ACS will have to employ several sensors installed in a manner that
will eliminate (or minimize) the probability that the onset of acoustic instability is
not detected. The observer or analyzer model of the ACS will also have to respond
nearly instantaneously with the desired control action and the controller will have
to respond, again in real time to allow the ACS to damp the instability before it
damages the combustor and/or adjacent system components. Some past studies
have shown that control of the most unstable mode can result in excitation of insta-
bility in other modes and thus, future ACS must have the capability to simultane-
ously control several unstable modes. Finally, to effectively damp various
instabilities, the ACS may require several actuators that will have to be optimally
installed within the combustor and could periodically modulate the fuel injection
rate with the large amplitudes required to damp the instability. Again, as in the case
of the sensors, improper placement of the actuators relative to the locations of the
combustion process heat release zone and/or the nodes and antinodes of the excited
acoustic oscillations may render them ineffective because they could potentially
require an inordinate amount of power to damp the instability. Finally, all of the
ACS components and their packaging will have to be sufficiently robust to survive
in a harsh engine environment over a long period of time.
Whereas most efforts aimed at actively controlling combustor processes to date
have studied ACS for controlling combustion instabilities and LBO, future
research may demonstrate control of other system critical parameters, e.g., emis-
sions, pattern factor, and engine health. Future ACS applications may also involve
detection and prevention of wall or component overheating, and may have capa-
bilities for monitoring and prognostication of combustor components’ health.
The expected operation of future intelligent combustors suggests that research
efforts will seek to develop “smart” sensors and wide-bandwidth fuel injectors
with multitasking capabilities. Future systems are likely to employ multi-
functional sensors and actuators shared by several ACS and the health monitoring
and prognostication system. For example, the same pressure sensors could be
possibly used to detect the onset of combustion instability and LBO, as well as
improper combustor operation. Furthermore, smart high-bandwidth fuel injec-
tors could be possibly used to damp combustion instabilities, compensate for loss
of combustion in a specific combustor region due to partial plugging of one of the
injectors, and to improve the pattern factor. Sophisticated analysis software and
COMBUSTION CONTROL 351

hardware to determine the state of the combustor in real time will be critical for
any ACS to be used in a real system. Finally, for practical use, future sensors,
actuators and computer software and hardware will have to be lightweight and
sufficiently robust to allow their long-term operation without failure in the harsh
combustor and engine environment.
The development of the advanced ACS will also require improved understand-
ing of the fundamental processes that control the performance of the combustor
when operated with and without control. Although experimental studies can
quickly demonstrate the viability of the control strategy and even deploy it in an
operational system, an in-depth understanding will require more detailed studies.
In this regard, numerical simulations (both LES and model-based studies) are
likely to play an increasingly important role as processing speed increases, and
simulations tools and models become more sophisticated and efficient. A major
advantage of simulations is that many properties that are difficult to measure or
observe in the experiments can be obtained in the entire domain with temporal
resolution of interest. However, predictive tools will have to be validated without
ad hoc model adjustments and for this to be accomplished, proper characterization
of the geometry and boundary conditions will be needed from the experimentalist.
Thus, coordinated experimental and numerical/modeling studies will be needed to
ensure that probability of future success in actual application is increased.

Acknowledgments
The authors would like to thank Mitat Birkan and Julian Tishkoff (AFOSR),
David Mann (ARO), Skip Fletcher (NASA), Dave Wisler and Hukam Mongia
(GE Engines), Anil Gulati (Siemens/Westinghouse) and Keith McManus (GE
GRC) for funding some of the reported results. Computational time was provided
by the Department of Defense High Performance Computing Modernization
Office at the HPC Centers at ERDC and NAVOCEANO, MS.
Chapter 11

Aeroacoustics of Flow Control

William Devenport*
Virginia Polytechnic Institute and State University, Blacksburg, Virginia
and
Stewart Glegg†
Florida Atlantic University, Dania Beach, Florida

I. Introduction
Active flow control (AFC) offers a number of exciting possibilities in aero-
dynamics, not just in reducing drag, increasing lift, enhancing propulsion and
reducing signatures, but also in freeing the designer from the one-to-one relation-
ship between a passive geometry and its aerodynamic performance. Aeroacoustics
has two roles to play here. Firstly, flow control can be and is used extensively as a
means to reduce flow generated noise. Secondly, flow control devices or the flows
they generate may themselves be significant sources of parasitic noise.
The simplest example in which a controlled flow creates sound is a musical
instrument such as a trumpet or a flute. The flow into the mouthpiece of the instru-
ment is carefully controlled by the lips of the musician, and the resonant cavity
which makes up the body of the trumpet or flute amplifies the sound generated in
the mouthpiece to create “wanted sound” or music. In contrast, the flow through a
valve or restriction in a piping system can cause “unwanted sound” or noise. There
are also examples in which sound is used to control a flow, the classic case being
the disturbance of a candle flame by a violin (LeConte, 1858; see also discussion
by Tyndall, 1881). While flow control by sound is not the topic of this chapter, we
will give a brief discussion of some of the more important effects of acoustic flow
control in aeronautical applications where appropriate.
Joslin et al. (2005) discuss in some depth the synergism between flow
control and noise control. They distinguish direct and indirect relationships. For

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Professor, Department of Aerospace and Ocean Engineering. Associate Fellow AIAA.
†Professor, Department of Ocean Engineering. Associate Fellow AIAA.

353
354 W. DEVENPORT AND S. GLEGG

example: blowing near the trailing edge (TE) of a fan blade to improve the TE
flow, thereby reducing drag but also reducing the noise produced by a following
stator (direct); active control of separation on a wing enabling a lower landing
speed and thus a reduction in noise (indirect). Connections of either kind can, of
course, be counterproductive as well—the fan blade blowing slot itself could (one
supposes) generate more noise than that saved at the stators.
In this chapter we discuss the basic principles of aeroacoustics and their appli-
cation to AFC through a number of examples that reference flow control strategies
and actuators. Our objective is to illustrate that there is often a need to consider the
aeroacoustic implications of flow control, whether or not the purpose of that
control is a reduction of flow-generated sound.
Our overview of aeroacoustic theory is intended to be a summary of the impor-
tant concepts. It is presented with the minimum of mathematics in a form that
aims to be accessible to readers familiar with aerodynamics but not acoustics.
Sufficient references are included so the reader can find detailed derivations where
necessary. In our discussion of applications, we consider AFC to be control that
involves energy input to the flow (such as through Coanda systems, plasma actua-
tors, fluidic jets, TE blowing) or involves actuation on a timescale of the order of
the timescales present in the flow (such as vibrating flaps, moving surfaces or, for
that matter, helicopter rotors).
In the following section we discuss the basic concepts of aeroacoustics and how
these apply to flow control devices. We then discuss applications of flow control
to noise reduction with reference to leading edge (LE) noise in fans and rotors
as well as TE noise sources, separated flows, and acoustic control of jets. Parasitic
noise from flow control devices is then discussed with reference to applications in
circulation control devices and synthetic jets.

II. Sound Generation by Flow


Sound consists of small perturbations of pressure that propagate through the
medium. The simplest way to generate sound is to use a vibrating surface such as
a loudspeaker diaphragm (Dowling and Ffowcs Williams, 1983). In this case the
motion of the surface causes time varying pressure fluctuations that propagate
over large distances as sound waves. This is an efficient acoustic source because
the displacement of the diaphragm causes an apparent net mass injection into the
fluid. The flow over an impenetrable surface or free turbulence in air generates
sound in a completely different way, which is far less efficient because there is no
net volume displacement of the fluid. The exception to this is an underwater
surface that causes cavitation bubbles, which collapse when they move into
regions of higher pressure or scatter pressure fluctuations caused by turbulent
flow. The bubbles can act as volume displacement sources and can dramatically
increase the radiated sound levels (Dowling and Ffowcs Williams, 1983).

A. Sound from Free Turbulence


Sound generation by flow was addressed in the seminal paper of Lighthill
(1952), using a model based on an acoustic analogy. The purpose of the analogy
is to evaluate the sound field in a fluid at rest caused by a source region of finite
volume that contains a turbulent flow. Lighthill argued that the acoustic waves in
AEROACOUSTICS OF FLOW CONTROL 355

the fluid outside the source region must propagate in accordance with the assump-
tions of linear theory, and could be described by the homogeneous wave equation.
However, the Navier–Stokes equations apply everywhere in the fluid and so, mak-
ing no assumptions, Lighthill rearranged these equations to obtain a wave equa-
tion for density perturbations in the stationary fluid surrounding the source region.
Specifically, by taking the time derivative of the continuity equation and subtract-
ing the divergence of the momentum equation, he was able to obtain the equation
(see Dowling and Ffowcs Williams, 1983)

∂ 2r¢ 2
2 ∂ r¢
∂ 2Tij
- co = (1)
∂t2 ∂ xi2 ∂ xi ∂ x j

where Tij = rvivj + pij - r¢c2odij and Einstein’s convention for the indices is fol-
lowed. For acoustic waves the density perturbation r¢ = r - ro (where r is the
density of the fluid and ro is the density of the stationary fluid surrounding the
source region) can be related to the pressure perturbation using the equation of
state p = r¢c2o and so the left-hand side of this equation describes the propagation
of acoustic pressure waves outside the region of turbulent flow. The right hand
side of this equation includes all the residual terms in the Navier–Stokes equations
and, if the assumptions of linear acoustics are applied, will be zero outside the
source region. Lighthill argued that the source term Tij described all the physics of
sound generation by flow.
The first term in the definition of Tij is rvivj where vi is the flow velocity. It
represents the instantaneous Reynolds stress contributions in the turbulent flow.
The second term is the compressive stress tensor pij, which includes the effects of
pressure perturbations in the turbulent flow and viscous stresses. For high Reynolds
number flows, the viscous stresses are often ignored, and so for flows in which the
relationship between pressure and density is uniform throughout the medium (for
example in underwater applications or low Mach number flows), we can assume
that pij - r¢c2odij = 0. This is not the case for all flows and for heated jets the second
term in Tij is considered to be the dominant source of sound (Morfey, 1973).
Consideration of Lighthill’s Eq. (1) shows that the analogy does not permit
sound waves to propagate at any speed other than the speed of sound in the
medium outside the source region. In reality, high-speed heated flows will have
significant variations in sound propagation speed within the source region, and so
sound propagation from a source element in the turbulent flow cannot be described
by the wave operator in Eq. (1). The analogy wraps up these effects in the source
term Tij but it is often argued that this is not correct. Alternative theories have been
derived (Howe, 1975; Lilley, 1974) which properly account for the propagation of
the sound waves through the mean flow, but these have not been applied as exten-
sively as Lighthill’s analogy.
The formal solution to Lighthill’s wave equation in the absence of any scatter-
ing surfaces gives the density perturbation outside the source region as

∂2 dV
r ¢(x, t )co2 = ÈT (y, t )˘˚
∂ xi ∂ x j ÚV Î ij
r = x-y (2)
t =t -r / co 4p r
356 W. DEVENPORT AND S. GLEGG

At large distances from the sources the double derivative outside the volume
integral can be approximated as (xi xj /c2or2)∂2/∂t2. This approximation can be made
because at large distances the field only contains sound waves propagating radi-
ally from the source region. Since these propagate at a fixed speed, the spatial
derivatives can be accurately written in terms time derivatives at a fixed point.
Turbulence with a lengthscale L and convection velocity U has a timescale of L/U
(see, for example, Tennekes and Lumley, 1972), and so the double derivative
outside the volume integral scales as (U/Lco)2. The volume integral of the Reynolds
stress term in Eq. (1) scales as U2L3 and so the rms density perturbation is expected
to scale as ro LU4/c2or. The far field acoustic intensity, which depends on the mean
square pressure perturbation, scales with the eighth power of the free-stream
velocity and the square of the lengthscale. This classic scaling law for the noise
from turbulence has been verified experimentally for cold jets (with the length-
scale proportional to the jet diameter), but not for heated flows for which the
Reynolds stress term in Eq. (1) is not dominant.
From the perspective of flow control it is important to recognize the correct
coupling between the scales of the turbulent flow and the propagating acoustic
waves. Some insight can be obtained by considering the density perturbation in
the frequency and wavenumber domain. Taking the four-fold Fourier transform of
Eq. (1) with respect to time and the three space dimensions gives

(w 2 - |k|2c2o )r¢(k, w) = ki kj Tij(w, k)

The inverse wavenumber transform then gives the density perturbation in the
frequency domain as
.
ki k j Tij (w , k )e - ik x
r ¢(x, w )co2 = Ú 2
dk (3)
k
(w /co )2 - k

The integrand is clearly dominated by its singular value which occurs when the
magnitude of the wavenumber of the turbulence matches the wavenumber of the
acoustic wave w/co, and requires that each wavenumber component ki is less than
w/co. In general, turbulence is dominated by eddies which are convected at the
mean flow speed and so most of the energy in the wavenumber spectrum is con-
centrated at the wavenumber (in the flow direction) of k1 ~ w/U. For subsonic
flows this is always greater than w/co. Most of our knowledge about turbulent flow
is based on our understanding of what happens at wavenumbers close to (or above)
the convective ridge where k1 ~ w /U. However, to understand or control the sound
radiation we need to address the low wavenumber regime which couples directly
to the acoustic waves. This has proven to be one of the most challenging problems
in aeroacoustics and must be given proper consideration when designing a flow
control device for jet noise.

B. Turbulence Close to Moving Surfaces


Lighthill’s theory can be extended to include situations where the flow encoun-
ters impermeable surfaces such as wings or propeller blades. This was first done
AEROACOUSTICS OF FLOW CONTROL 357

by Curle (1955) for stationary surfaces and later extended to moving surfaces by
Ffowcs Williams and Hawkings (1969). When moving rigid surfaces are present
the acoustic field is given by the FW–H equation [see Dowling and Ffowcs
Williams (1983)]

∂2 È Tij (y,t ) ˘ ∂ È pij (y,t )n j ˘


r ¢(x, t )co2 = Ú ÍÍÎ 4p r 1 - M ˙ dV (y ) - Ú ÍÍÎ 4p r 1 - M ˙ dS (y )
∂xi ∂x j r ˙
∂ xi r ˙
Vo ˚t =t * So ˚t =t *
∂ È roV j (y,t )n j ˘
+ ÚÍ
∂t Í 4p r 1 - Mr ˙
S Î
˙
˚t =t *
dS (y ) (4)
o

where t * = t - r/co. In this equation the first term represents the sound generated
by the flow and differs from Eq. (2) because the volume (and hence the y coordi-
nate) moves with the surface, for example a propeller blade. The term Mr is the
component of the source Mach number in the direction of the observer, and 1 - Mr
is defined as the Doppler factor. In addition to the noise generated by the flow
there are contributions from the compressive stress tensor on the surface, and the
surface velocity Vj.
The last term in Eq. (4) is often referred to as thickness noise because it is
completely determined by the size and velocity of the body. It is zero unless the
surface is accelerating, and can be a large contributor to the sound from transonic
propeller blades. Notice how it only depends on the density in the surrounding
medium, not the local density on the surface. However, if an actuator, which
causes a mass displacement, is added to the surface, then an additional term ruj
must be added to the thickness noise term where r is the local density and uj is the
velocity of the actuator relative to the surface. An actuator of this type can be a
very efficient acoustic source and cause sound to radiate to the acoustic far field.
Introducing mass displacement actuators may suppress turbulent flow, which will
reduce the first two terms in Eq. (4), but the mass displacement by the actuator
directly couples with the acoustic field and increased sound radiation may occur.

C. Acoustically Compact Surfaces


The second term in Eq. (4) is often referred to as loading noise and depends on
the compressive stress tensor pij, which includes both the effect of local pressure
and the viscous stresses at the surface. The latter are usually negligible and surface
pressure fluctuations dominate. For objects that are small compared to the acoustic
wavelength the surface sources can be treated as a point source whose strength is
equal to the integrated surface pressure, or equivalently the unsteady loading F(t).
The scaling of this source with flow speed is determined in the same way as above
by using the far field approximation for the derivative outside the integrand
∂/∂xi ~ -(xi/r)(1/co)∂/∂t, giving

xi ∂Fi (t - r /co )
p ( x, t ) = (5)
4p r 2 co ∂t
358 W. DEVENPORT AND S. GLEGG

It is important to note that this source type depends on the net unsteady force
applied to the fluid by the surface. This implies that in flow control applications
the radiated sound from an acoustically compact body will be reduced if a flow
control device can be developed that reduces the net unsteady loading on the sur-
face. Also note that this type of source is directional, and that (xi /r)∂Fi /∂t represents
the rate of change of force resolved in the direction of the observer. The acoustic
field is therefore a maximum in the direction of the force, and minimal at 90 deg
to the direction of the force. Because time scales on flow speed U and lengthscale
L as L/U and the force as rU 2L2, the acoustic pressure from unsteady loading noise
will scale as p ~ rU3L/rco, and the far field intensity will increase as U6. If we
compare loading noise with the noise from free turbulence we find that the acoustic
intensity of the turbulence noise is of order M2 less than the loading term. For low
Mach number flows with no mass displacement source, the loading sources will
therefore dominate. Furthermore, noise reductions using a flow control device are
possible if the unsteady loading is reduced. To estimate the effect this may have for
compact bodies we consider the power spectrum of the acoustic field Spp in terms
of the spectrum of the net unsteady loading on the surface SFF, as
2
Ê w cos q ˆ
S pp (x, w ) = Á S (w ) (6)
Ë 4p co r ˜¯ FF

where q is the angle between the direction of the force and the observer. Modifying
the spectrum of the unsteady loading will therefore alter the radiated sound field.
If an additional mass displacement source is introduced as a flow control device
we can estimate its contribution from the third term in Eq. (4), which gives for
acoustically compact bodies:

ro ∂Q(t - r /co )
p(x, t ) =
4p r ∂t Ú
Q(t ) = u j (y, t )n j dS (y)
S
(7)

The actuator will be a significant new source of sound if its integrated surface
velocity Q is large compared to F/roco. This is more likely to occur in underwater
applications where the speed of sound is large and the flow speeds are slower than
in aeronautical applications.
It is also important to appreciate that the far field approximation for the deriva-
tive ∂/∂xi ~ (xi/r)(1/co)∂/∂t includes a direction cosine xi /r which defines the
directionality of the source. The sound radiation from a compact loading source
has a cosine directionality, which is aligned in the direction of the force. In con-
trast, the mass displacement term is omnidirectional, and so will radiate to parts
of the acoustic field where the loading noise is small. The acoustic radiation from
a surface actuator cannot be used to cancel loading noise in all directions unless
the actuator surface velocity distribution is carefully chosen to match the direc-
tionality of the loading source.

D. Large Surfaces
When the size of the surface becomes large compared to the acoustic wave-
length the properties of the loading noise can change significantly. For example,
AEROACOUSTICS OF FLOW CONTROL 359

if we consider a turbulent boundary layer on an infinite flat plate the surface pres-
sure fluctuations are caused by eddies that are convected at a speed slightly less
than the mean flow speed. The dominant wavenumber contributing to the pressure
fluctuation is of order w/U which is much larger than the acoustic wavenumber
w /co. Consequently the boundary-layer pressure fluctuations will be a very weak
source of sound because their characteristic dimensions are very much smaller
than the acoustic wavelength in low Mach number flows. The situation changes
dramatically if the boundary layer encounters a discontinuity such as a step or a
gap. The sudden change in the flow at a discontinuity will cause a local hot spot
in the surface pressure fluctuations, which is fixed to the surface and not convected
by the flow. This fixed disturbance causes a net unsteady loading on the surface
that acts as an acoustic source. The example given above refers to the boundary
layer interacting with a discontinuity but similar principles apply to a boundary
layer convecting past a TE (TE noise), or a gust interacting with the LE of a blade
(LE noise). In this case a compressible pressure fluctuation is scattered by the
edge so that the turbulent flow matches the boundary conditions on the plate as
well as the requirements for homogeneous flow upstream and downstream of the
edge. If the surfaces are very large, and can be considered semi-infinite, the scat-
tering mechanism at the edge has a frequency dependence which scales as
(w /co)-1/2 and so by using the scaling laws defined above we can show that these
sources scale with the fifth power of the flow speed.
In this section we have introduced some of the fundamental concepts of aero-
and hydroacoustics and in the following sections we will consider the applica-
tion of these concepts to some flows of practical interest.

III. Leading Edge Noise


One of the primary sources of sound generation by propellers, helicopter rotors,
and fans is the interaction of the blades with flow disturbances. Examples include
blade vortex interactions on helicopter rotors, blade wakes interacting with down-
stream rotors or stators in turbomachinery, and free-stream turbulence ingested
into automotive cooling fans. In all cases the sound is generated by the unsteady
surface pressure on the blade surfaces, and typically scales with the sixth power
of the blade speed. The most effective way to reduce the noise from fans and
rotors is to reduce the tip speed, but in special cases there are alternative approaches
which can be used to minimize the magnitude of the gust which the blade encoun-
ters, and hence reduce the unsteady loading and the radiated noise.

A. Blade Vortex Interactions


As an example we will consider a blade vortex interaction on a helicopter rotor.
In certain flight regimes such as a slow descent, a helicopter rotor ingests its own
wake and the blades slice through the trailing vortices shed by the tips of the rotor
blades. These vortices have very intense local velocities that cause a rapid fluctua-
tion in angle of attack on a rotor blade that passes close by. The interaction is most
dramatic when the axis of the vortex is parallel to the LE of the blade so the entire
span of the blade encounters the gust at the same time. If the axis of the vortex is
skewed or displaced relative to the blade LE then the unsteady loading and noise
is reduced. The basic principle of noise control in helicopters is to avoid parallel
360 W. DEVENPORT AND S. GLEGG

blade vortex interactions wherever possible, but in many applications this cannot
always be achieved.
The unsteady loading caused by a blade vortex interaction is a function of blade
angle of attack. Figure 1 shows the unsteady loading as a function of the nondi-
mensional time Ut/c where c is the blade chord for a parallel blade vortex interac-
tion in incompressible flow. The vortex passes the LE of the blade at nondimensional
time of -0.5 and the trailing edge at nondimensional time of 0.5. The pulse in the
time history is caused by the interaction of the vortex with the LE of the blade, and
its magnitude increases with blade angle of attack. Also note that there is no pulse
in the unsteady loading as the vortex passes the TE. This would not be the case if
the unsteady Kutta condition were not applied, and so a control surface which
impacts the natural cancellation of the unsteady surface pressure at the TE could
generate extra sound.
The importance of blade vortex interactions on helicopter rotors has led to the
implementation of active blade control to reduce the significance of the interac-
tion. These systems are implemented using higher harmonic control of the blade
angle of attack. Harmonic control is required to ensure that a helicopter is stable
in forward flight, reducing the angle of attack on the advancing side and increas-
ing it on the retreating side. Higher harmonic control is a modification to the basic
system that reduces the angle of attack during a blade vortex interaction or
increases the distance between the blade and the vortex during critical maneuvers.
Brooks et al. (1994) describe the results of a wind tunnel test using higher harmonic
control in which the blade vortex interaction noise was reduced by 6 dB for blades
on the advancing side of the rotor and similar amounts for retreating blades, depend-
ing on the phase of the higher harmonic control. More recently, Nguyen et al. (2000)

1.5

LIVE GRAPH
Click here to view
1
CL(t)

0.5

–0.5
–1 –0.5 0 0.5 1
Ut/c

Fig. 1 The unsteady lift CL(t) = L(t)/ρoUΓ during a 2-D blade vortex interaction with a
vortex of strength Γ, for an airfoil with a thickness to: chord ratio of 0.12 in incompress-
ible flow. The three curves show the effect of blade angle of attack for 0 deg (solid line),
6 deg (dashed line), and 12 deg (dashed dot line). The vortex passes the LE at Ut/c = −0.5
and the TE at Ut/c = 0.5 and is displaced from the stagnation streamline by 0.1 c.
AEROACOUSTICS OF FLOW CONTROL 361

have demonstrated a 12 dB noise reduction on a full-scale XV-15 rotor demon-


strating its application to tilt rotor aircraft. At full scale a noise reduction of 3.5
EPNdB was demonstrated using higher harmonic control on a Gazelle helicopter
(Polychroniadis, 1990).

B. Rotor Stator Interactions


Another application of flow control associated with LE noise is in turbomachin-
ery where a stator is sited downstream of a rotor. The wakes from the rotor blades
have a mean velocity deficit and contain turbulence. The mean velocity deficit is
nearly identical for each rotor blade wake and so, as the wakes wash over the
stators, each stator sees a periodic gust. The resulting unsteady loading on the
stator vanes results in sound radiation at multiples of the rotor blade passing
frequency, referred to as rotor–stator interaction tones. Superimposed on this
periodic sound field is a random broadband sound field generated by the aperiodic
impact of the turbulent structures contained in the rotor blade wakes. The same
mechanisms occur in designs where the wakes of fixed inlet guide vanes wash
over a moving rotor.
The tone-noise component of the rotor stator interaction is a particularly appeal-
ing target for active control because of its deterministic nature and so a variety of
different approaches have been attempted or proposed. One approach (which may
well fall outside the scope of active flow control) is the direct generation of a
canceling sound field, such as using actuators embedded in the inlet of the turbo-
machine (Thomas et al., 1993) or on the surface of the downstream stator vanes
themselves (Kousen and Verdon, 1994; Sawyer and Fleeter, 2000; Sawyer et al.,
1997; Schulten, 2001). Note that the goal of this latter method is the cancellation
of the radiated acoustic part of the unsteady stator pressure field, not of the entire
unsteady lift. However, whatever the method, the acoustic power required to
cancel tones in, say, an aircraft engine implies the use of actuators that maybe both
be large and heavy.
A second, closely related approach that avoids this problem is to use flow con-
trol to generate the canceling sound field indirectly. This possibility was first
realized by Nelson (2000) who proposed installing wake generators on the casing
upstream of a fan so as to generate a secondary sound field to cancel that produced
by inflow distortion. Such a system also lends itself to closed loop control (Kota
and Wright, 2006) where the wake generators are adjusted in order to cancel an
unknown flow distortion. Polacsek and Desbois-Lavergne (2003) implemented an
open-loop system of this type for the control of rotor–stator tones in a research
compressor. A set of radial rods, mounted on a rotating ring, were placed in front
of a rotor and following stator. Rotation allowed the phase of the secondary sound
field to be adjusted reducing the intensity of stator–rotor tone at the blade passing
frequency by 8 dB. Perhaps surprisingly, the predicted effects of the rods on the
aerodynamic performance of the compressor were found to be negligible. A
somewhat different method of generating a secondary sound field is to introduce
disturbances downstream of the rotor. Ashcroft and Schultz (2004) report that jets
injected radially through the casing downstream of the rotor can reduce tone noise
levels by as much as 20 dB. They show computationally that the canceling sound
field originates from the interaction of the rotor blades with the upstream (potential)
362 W. DEVENPORT AND S. GLEGG

disturbance produced by the jets. This implies an exponential decay in the cancel-
ing sound field as the distance between the rotor and jets is increased. An advan-
tage of introducing disturbances downstream of the rotor is that it reduces their
potential to serve as sources of additional broadband noise.
Other proposals for eliminating the source or generation of stator–rotor interac-
tion noise include: actuated flaps on the stator TEs driven so as to reduce the
unsteady lift produced by interaction with an oncoming wake (Simonich et al.,
1993); LE actuators on the stators (Kerschen and Reba, 1995); TE flaps on inlet
guide vanes (ahead of a rotor) designed to move propulsively and thus cancel the
wake deficit (Opila et al., 2004); and the use of TE blowing to cancel the wakes of
an upstream rotor or vane set. Of these technologies, TE blowing has received the
most attention and appears the most promising. Furthermore, this appears to be
the only AFC method that offers any promise of reducing broadband interaction
noise as well as tones.
Much of the recent interest in the use of TE blowing began with the work of Sell
(1997) and Brookfield (1997), also reported by Brookfield and Waitz (2000).
Using a linear cascade, Sell established that blowing through holes in the TEs of
compressor blades could both “fill” the wake and lower the turbulence levels
within it, producing a momentumless wake for a blowing rate equivalent to 1% of
the total mass flux through the blade row. Brookfield confirmed these exact same
effects on a model fan modified to allow blowing through an array of holes drilled
into the TEs of the fan blades. Blowing rates equivalent to about 2% of the total
mass flow through the fan not only flattened the fan blade wake profiles and
reduced turbulence levels within them, but also greatly reduced unsteady pres-
sures on the downstream stators away from the hub. Similar rates of blowing and
reductions on unsteady forces on downstream stators were observed by Wo et al.
(2002) in a low-speed fan, who used injection through part-span slots located near
the midspan of the blade TEs. Borgoltz et al. (2005, 2006) showed that blowing
also has dramatic effects on the eddy structure of the wake downstream of com-
pressor blades. Blowing rates well below those needed to flatten the wake are
sufficient to significantly reduce the correlation scales of wake eddies, as well as
their intensity. Borgoltz et al. (2006) estimates that these combined effects could
reduce broadband noise generated by a downstream stator by as much as 7 dB.
Actual indications that suppression of the wake through TE blowing reduces
sound levels exists in the form of acoustic measurements on a fan with blown inlet
guide vanes (Leitch et al., 2000; Rao et al., 2001) and on the Active Noise Control
model fan at NASA Glenn by Sutliff et al. (2002). They showed convincingly that
blowing rates of 1.6–1.8% of the total mass flux through TE slots in the fan blades
lower tone noise levels from the inlet and exhaust by as much as 19 and 12 dB in
the far field. Sutliff (2005) looked at broadband effects and found substantial
reductions in fan wake turbulence levels and unsteady stator blade pressures con-
sistent with reduction in broadband noise. This could not be confirmed directly,
however, since broadband noise from the ANCF is dominated by noise from the
rotor alone rather than rotor–stator interaction. The magnitude of any realizable
benefit therefore remains an open issue, and there have been no comparative
studies that balance this against noise from the injection itself.
TE blowing has not yet seen practical application in aircraft engines, nor do
such applications appear to be imminent. The primary reason is the amount of
AEROACOUSTICS OF FLOW CONTROL 363

blowing required and the complex systems required to supply it. A blowing rate of
2% in the bypass of a commercial aircraft engine would require diversion of
10–20% of core mass flux, depending on bypass ratio. This is more than can be
tolerated without an overall redesign of the engine which, of course, might have
larger indirect detrimental effect on noise.
It seems well established that blowing rates are dependent on the method of
injection. Blowing at low rates through a slot on the suction side of the TE can
actually deepen the wake (Borgoltz et al., 2005). Injecting air in jets on both sides
of the blade at the 80% chord location appears to require the least mass flux
(Langford et al., 2005). This sensitivity is a clear indication that benefits in the
wake are a combination of improving the flow past the TE to minimize any low-
momentum regions and canceling the wake momentum deficit. Halasz et al.
(2005) show that mass flow rates can be further reduced by applying TE blowing
to only a subset of the fan blades. While this is not as effective in reducing all the
tones, this can be compensated for in the acoustic liner of the engine. The net
effect is a predicted reduction of about 10 dB in tone noise when blowing about
0.9% of the overall engine mass flow.

IV. Trailing Edge Noise


TE noise is one of the most important sources of broadband noise from fans and
plays a significant role in airframe noise. It is caused by the interaction of the
turbulent boundary layer on the surface of the blade or wing interacting with the
TE. This source mechanism has been studied in detail by both Howe (1998) and
Brooks et al. (1989) who have shown that the source levels depend critically on
the properties of the boundary layer at the TE. A model for the characteristic
spectrum of TE noise is given by Howe (1998) in terms of the power spectrum of
the far field acoustic pressure generated by a turbulent boundary layer on a semi-
infinite flat plate. The noise is generated as the boundary-layer turbulence passes
over the edge, and far field power spectrum is

sin 2 (f /2)sin y
S pp (x, w ) = Uc L dSPP (w ) (8)
2p 2 r 2 co

where f is the angle of the observer to the direction of the flow, y is the angle of
the observer to the direction of the TE, Uc is the convection velocity of turbulence
in the boundary layer, d is the blade span, and L the spanwise correlation length
scale of the boundary-layer pressure fluctuations. The function SPP is the wall
pressure spectrum beneath the boundary layer well upstream of the TE, and is
expected to scale as ( rov*2)2d */Uc. The TE noise levels are therefore directly pro-
portional to the boundary layer displacement thickness in this model.
Figure 2 shows spectra of TE noise measured by Brooks et al. (1989b) in an
open jet windtunnel as a function of blade angle of attack. As the angle of attack
is increased the boundary layer thickens at the TE and the noise level increases.
From a flow control perspective, circulation control (see following section) which
will increase lift while minimizing the boundary layer thickness at the TE should
reduce noise. However this does not take into consideration the noise generated
364 W. DEVENPORT AND S. GLEGG

80

LIVE GRAPH 75
Click here to view
70 3deg
1.5deg
1/3rd Oct SPL

65

60
0deg
55

50

45

40
100 101
Frequency in kHz

Fig. 2 The effect of angle of attack on TE noise. Measurements by Brooks et al. on a


NACA 0012 airfoil with a tripped boundary layer. Flow speed 71.3 m/s; chord 0.3048 m.

by the flow control device itself, or the details of the mean flow modification that
will vary from case to case.
TE noise is also affected by the transition point of the boundary layer on the LE
of the blade. Figure 3 shows the radiated noise measured by Brooks et al. on a
blade with different boundary layer trips. The noise level increases, and has a
lower peak frequency, as the size of the trip is increased which is consistent with
the thickening of the boundary layer at the TE. Laminar flow control, which delays

80

75
LIVE GRAPH
Click here to view 70
Tripped Slightly tripped
1/3rd Oct SPL

65

60

55 Untripped

50

45

40
100 101
Frequency in kHz

Fig. 3 The effect of boundary layer trips on TE noise. Measurements by Brooks et al.
on a NACA 0012 airfoil. Flow speed 71.3 m/s; chord 0.3048 m; 0 deg angle of attack.
AEROACOUSTICS OF FLOW CONTROL 365

transition, should reduce TE noise providing the properties of the boundary layer
at the TE are not adversely affected.
For low Reynolds number flows over airfoils, or flows over airfoils with blunt
TEs, coherent vortex shedding can occur at the TE. This can often result in a
discrete tone being generated at the frequency of vortex shedding, and is some-
times referred to as blade singing. A detailed discussion of these phenomena is
given by Blake (1986), and will not be discussed in detail here. However, either
active or passive methods for eliminating the coherent vortex shedding always
reduce the radiated noise, and flow control devices which achieve this effect may
be beneficial.

V. Separated Flows
When lifting surfaces operate at high angles of attack, the flow separates and
the source mechanism changes from TE scattering to unsteady loading caused by
the increased contribution of the fluctuating lift and drag forces. For example,
Fig. 4 shows the spectra measured by Brooks et al. (1989) for an airfoil as it goes
through stall. Note how the levels increase and the spectrum shape is dramatically
altered, giving much higher levels at low frequencies. Clearly this is a situation
where flow control can be used to eliminate separation and will have a positive
impact on the radiated noise, changing the source mechanism back to TE noise
that depends on the boundary layer features rather than the separated flow.

VI. Acoustic Control of Jets


A substantial amount of research has been done on the acoustic control of
turbulent jets (see Ginevsky et al., 2004 for a complete review). For low Mach

80
LIVE GRAPH 75
Click here to view Separated Flow
70
1/3rd Oct SPL

65

60

55

50 Attached Flow

45

40
100 101
Frequency in kHz

Fig. 4 The effect of flow separation on airfoil self noise. Measurements by Brooks
et al. on a NACA 0012 airfoil with a tripped boundary layer. Flow speed 71.3 m/s;
chord 0.1016 m; 0 deg and 8.9 deg angle of attack.
366 W. DEVENPORT AND S. GLEGG

number jets it is found that far field broadband noise is increased if the jet is
excited by a low-frequency acoustic actuator. In contrast the radiated broadband
noise is reduced if the jet is excited by a high-frequency acoustic source. However,
if the jet exit velocity is transonic, the high-frequency attenuation is not achieved
(Ginevsky et al., 2004). These effects are explained by the presence of large-scale
coherent structures in the jet in which small-scale turbulence is embedded. The
low-frequency acoustic excitation acts directly on the large-scale structures and
alters the rate of mixing in the jet, and hence the radiated sound. More details on
this subject can be found in the detailed text on the acoustic control of jets by
Ginevsky et al. (2004).

VII. Circulation Control


An example of flow control devices affecting TE and separated flow noise
sources can be found in the active control of circulation using the Coanda effect.
Here a wall jet is injected tangential to the suction surface of an airfoil a short
distance upstream of the TE of an airfoil (Fig. 5). The TE itself may be rounded
so that the additional momentum added by the jet carries the flow around the TE
before it detaches from the pressure side. The detachment point, and thus the cir-
culation and lift on the airfoil, are determined by the strength of the jet. An alter-
native arrangement, more compatible with cruise configurations, is to use the
Coanda jet to drive the flow around a small hinged flap at the TE deflected at a
high angle. Circulation control also has application to thrust vectoring, above-
wing engine configurations where the engine exhaust serves as the jet, and (for
pulsed blowing) to control of lifting surface vibration (Raghavan et al., 1988).
This basic flow arrangement of Fig. 5 allows for a number of potential noise
sources. There is the quadrupole noise generated by turbulent motions within the
Coanda jet itself (source 1) and then there are a number of TE sources. First, tur-
bulence in the airfoil boundary layer scatters as this boundary layer passes over
the slot lip (source 2). Second, turbulent fluctuations in the Coanda jet exiting the
slot will scatter from the same edge (source 3). Finally, the rounded TE may be a
source of scattering for turbulence in the combined boundary layer and jet that
flows around it (source 4).

Slot
Suction side Jet
boundary Rounded trailing edge
layer 2
3 1

Fig. 5 Noise sources associated with a Coanda circulation control device at the TE of
a lifting surface. Numbers indicate locations of sources referred to in text.
AEROACOUSTICS OF FLOW CONTROL 367

Howe (2002) considers the noise generated by sources 2–4 in the context of a
Coanda effect hydrofoil where jet noise would be negligible. He is able to develop
models for each of these sources by using conformal mapping to generate the neces-
sary Green’s functions and by expressing the source in terms of the wall pressure
spectrum of the boundary layer (or slot flow) as it approaches the scattering edge.
Howe develops equations for predicting the noise contributed by each of these
sources. Because the mechanism of noise production is essentially the same, the
functional form of the noise spectra is quite similar, differing only by weighting
functions related to the local geometry. All indicate a variation of overall noise
levels with the 5th power of velocity and scaling of frequency on velocity and
displacement thickness. These scales are, more specifically, the convection veloc-
ity of the turbulence that is being scattered, and the displacement thickness of the
layer in which it resides, and are thus different for each of the sources.
Howe performs calculations based on these formulae for the case of a hydrofoil
with a sharp-edged slot 0.3% of the airfoil chord in height and a jet exit velocity
equivalent to six times the free-stream, and compares them with predictions of TE
noise for a hydrofoil with sharp TE and no circulation control. He shows that
contributions from the rounded TE increase noise levels at reduced frequencies
(angular frequency normalized on semichord and flow velocity) wr below 100 by
about 5 dB. A 1–2 dB increase is produced by the external boundary layer inter-
acting with the slot lip for 100 < wr < 1000, whereas turbulence in the Coanda jet
increases levels by as much as 30 dB for wr > 1000. The large influence of sound
scattered from the Coanda jet is the result of the high velocities there, and illus-
trates the importance of arranging for the jet to be laminar (and thus silent) if
possible. Howe also demonstrates that using a blunt slot lip substantially reduces
scattering effectiveness and thus the noise levels it generates.
The jet noise component ignored in Howe’s analysis is an important, if not
dominant, component at the higher Mach numbers of many aerodynamic applica-
tions. Lighthill’s theory of jet noise (see above) predicts that the sound intensity
radiated by a jet should vary as the 8th power of the jet velocity U and the square
of the jet diameter D. These predictions have largely been borne out by experi-
ments. Almost all these experiments, however, have been carried out on circular
or non-circular jets with small aspect ratios compared to those that might be
expected in a practical Coanda system.
One exception is the work of Munro and Ahuja (2003) who measured noise
radiated from rectangular jets with aspect ratios 100–3000 using an array of
microphones arranged along an arc perpendicular to the plane of the jet for sub-
sonic jet speeds between 500 and 1100 ft/s. They confirm that for a jet with the
proportions of a Coanda system the sound intensity still varies as U8 but the direc-
tivity is significantly different from a round jet with levels much lower close to the
jet axis. Furthermore, Munro and Ahuja show convincingly that noise levels are
not correlated by the equivalent jet diameter (the geometric average of the height
h and width w) used in prior studies of non-circular jets. Instead their data indicate
a length scale proportional to h3/2w1/2. Munro and Ahuja do not offer a physical
explanation for this scaling.
Salikuddin et al. (1987) performed acoustic measurements on a circulation con-
trol wing with a chord of 38.1 cm, and a rounded TE with a radius of some 2% of the
chord. Air was exhausted from a slot ahead of the TE with an adjustable height
368 W. DEVENPORT AND S. GLEGG

(between 0.3 and 1.22 mm) at speeds between 150 and 340 m/s. Microphone mea-
surements were made in an arc in the center-span plane 2.44 m from the TE. These
measurements show the noise increasing with the jet velocity to the power of 7.34
and 7.5. It therefore appears likely, in the light of Munro and Ahuja’s (2003) study,
that jet noise was the dominant (but not sole) contributor. It is interesting that, while
not noted at the time, Salikuddin et al.’s (1987) measurements show noise levels that
increase approximately with the three-halves power of the slot height.
For supersonic Coanda jet velocities it is likely that additional noise-producing
mechanisms may appear. On an axisymmetric Coanda configuration (with rele-
vance to industrial gas flares) Carpenter and Green (1997) observed tone noise
resulting possibly from jet screech (feedback from the wavefield of the jet,
reflected by shock cells, to the instabilities in its mixing layer just downstream of
the slot lip) and from instability in the Coanda nozzle for pressure ratios just large
enough to produce choked flow. An additional mechanism they propose, appar-
ently unique to Coanda systems, is feedback to the mixing layer at the lip from the
intense acoustic field formed around the reattachment point of a separation bubble
located on the downstream curved surface. Interestingly, Carpenter and Smith
(1997) found that using a sawtooth profile at the slot exit eliminated the tone noise
sources by introducing streamwise vorticity—a solution somewhat reminiscent of
the use of chevrons on jet engine exhausts.
While there is no question that there are many opportunities for noise genera-
tion in Coanda systems the important issue aeroacoustically is whether they are
less noisy than alternatives that produce the same aerodynamic result. This issue
is directly addressed in the work of Munro et al. (2001) who compared the noise
produced by a supercritical airfoil with circulation control and with a single TE
flap at approximate equal lift conditions. Circulation control was achieved using a
Coanda jet blowing over a small deflectable flap whose angle was adjusted
between 30 deg and 90 deg. Blowing slot Mach numbers from 0.3 to 1.2 were
used with free-stream speeds from 100 to 250 ft/s. The airfoil chord was some
8 in. and the slot height was set from 0.003 to 0.02 in. They found that the major-
ity of the noise generated came from the jet, and noise levels varied as the 8th
power of the velocity, except for noise generated internally by the air supply sys-
tem, and by tone noise generated by regular shedding from the back of the blown
flap when it was deflected at 90 deg (consequently 30 deg was used for most of
their tests). They found that, for the same lift enhancement, jet noise could be
reduced by increasing the slot height and reducing the jet velocity while maintain-
ing its mass flux. Comparisons with the flapped wing showed little aeroacoustic
benefit for the optimum blowing case, but Munro and co-workers point out that in
practical applications cutouts in the conventional flap would be required that
would not be necessary in the active controlled system. Tests with a cutout do
indeed show benefits to the blowing system for some conditions. Furthermore, the
implementation of the Coanda system would likely lead to indirect acoustic ben-
efits through the weight saved by not having a flap system.

VIII. Synthetic Jets


A particularly important AFC application is the use of a Helmholtz resonator as
a synthetic jet actuator. A typical actuator (Fig. 6) might consist of an orifice in a
AEROACOUSTICS OF FLOW CONTROL 369

Orifice

Cavity

Oscillating diaphragm

Fig. 6 Schematic of a synthetic jet actuator.

flow surface backed by a cavity. The cavity is driven by an oscillating diaphragm


(such as a piezoelectric element, a piston, or loudspeaker) so as to generate an
oscillating flow through the orifice. Because of the viscous asymmetry of suction
and blowing, under certain conditions (Holman et al., 2005) a series of vortex
rings is shed from the orifice that synthesizes a jet flow which is supplied with
fluid drawn in radially from around the orifice. Synthetic jets can be quite power-
ful, generating velocities of up to 250 m/s (Muller et al., 2001; Shaw et al., 2006).
They can also be microfabricated (Muller et al., 2001). Synthetic jets may be used
to deflect streamlines near the surface of an aerodynamic body and thus provide
active control of its effective shape (Rampunggoon and Mittal, 2002). In such
applications the driving frequency of the actuator is usually set to be much higher
than any natural frequencies in the flow. Another approach is to use the unsteady
blowing provided by the jet to couple with natural flow instabilities and therefore
amplify the control effect. Such a strategy has been used, for example in separa-
tion control (Greenblatt and Wygnanski, 2000). The reader is referred to other
sections of this book, to reviews such as Glezer and Amitay (2002), and to the
broad literature available for a discussion of the many fluid dynamic aspects and
applications of synthetic jets.
In discussing the aeroacoustics implications of synthetic jets we first consider
them as acoustic sources. The actuator itself causes a mass displacement and
behaves as a monopole source [specified by Eq. (7)] that can be a very efficient
acoustic radiator. Furthermore, synthetic jet actuators are usually driven at or near
resonance in order to maximize the mass displacement for maximum effective-
ness, but this also maximizes the sound they produce. Subjectively, these devices
are loud, some to the point of requiring hearing protection and this has the poten-
tial to be a limiting factor in applications where other noise levels are low, such as
in providing air flow to cool computer chips (see Seeley et al., 2006, for example).
To evaluate the sound radiation from synthetic jets we will consider them as small
mass displacement sources. If the exit flow velocity from the orifice is known,
then Eq. (7) can be used to calculate the radiated sound level. However, in some
cases only the pressure inside the cavity is known and so the jet should be modeled
370 W. DEVENPORT AND S. GLEGG

by a fluctuating pressure driving fluid through an orifice in a large plate. Howe


(1998) addresses the sound radiated through such an orifice subject to an incident
pressure fluctuation. He shows how the sound wave radiated from the orifice
becomes distorted as the amplitude of the fluctuating pressure in the cavity is
increased. At low pressure fluctuation amplitudes (around 10-4 atm or less) the
sound radiation is at the frequency of the resonator and undistorted. As the pres-
sure amplitude increases, however, harmonics start to appear in the radiated field
and a phase lag develops as a consequence of the nonlinear differential equation
governing the flow velocity in the orifice. At large amplitudes (around 10-1 atm)
these effects are quite marked. Howe’s theory is not complete—it specifically
neglects rotational flow effects and so the effect of vortex formation is not included,
which could be important. Crossflow effects are also left out. Lumped parameter
modeling of synthetic jet actuators has been performed by Gallas et al. (2003) that
includes (as an impedance) the acoustic radiation effect as well as the diaphragm
dynamics on the mass flow through the orifice. Acoustic calculations have also
been attempted by Seeley et al. (2006). However, a comprehensive theoretical
formulation that incorporates the effects neglected by Howe is clearly needed.
For high-speed synthetic jets quadrupole jet noise may also be a significant
broadband source, although one that may be mitigated by crossflow. While there
appear to be no measurements or theoretical work in this area, it seems likely that
there will be significant differences with conventional jets. Synthetic jets have
different growth and entrainment characteristics. Furthermore they display unusual
turbulence characteristics, including an absence of pairing in the shed vortices and
a rapid decay of high-frequency fluctuations indicating a high dissipation rate
(Smith and Glezer, 1998).
The indirect effects of synthetic jets on noise generation are of both concern
and interest in an aeroacoustic context. The action of a synthetic jet on an attached
turbulent flow will likely include an increase in boundary layer thickness and
turbulence and thus increase noise that might be radiated from a downstream TE
(Tian et al., 2006b). In preventing airfoil separation, a synthetic jet will exchange
separation noise for TE noise, but whether this increases or reduces the radiated
noise will depend on the application.
One approach to reducing the noise from synthetic jets is to use multiple jets
which are driven in antiphase. The net mass displacement from the sum of the two
jets is zero and so the monopole source described by Eq. (7) has zero source
strength. However this is an oversimplification and the sound level will not be
identically zero. Consider, for example, the combination of two single frequency
synthetic jets which have volume velocities Q1(t) = Q exp(-iwt) and Q2(t) =
-Q exp(-iwt). If the jets are located at d/2 and -d/2 then the far field is given by
the sum of the contributions the two jets. Using Eq. (7) gives

- iw ( t - x - d /2 / co ) - iw ( t - x + d /2 / co )
-iwroQe -iwroQe
p ( x, t ) = - (9)
4p x - d/2 4p x + d/2

Equation (9) shows that there will be a far-field sound generated by the two
jets in antiphase, whose strength depends on the differences in propagation dis-
tance from each source to the observer. The far-field sound can be obtained by
approximating the propagation distance |x ± d/2| as |x| - (±x . d/2|x|) which is a
AEROACOUSTICS OF FLOW CONTROL 371

good approximation when |x|  |d|. (Note: x . d is the dot product of the vector
x and vector d, and |x| - x . d/|x| = |x| - |d| cos (q), where q is the angle of the
observer to the line between the two sources.) Hence the radiated sound field is
given (see Dowling and Ffowcs Williams (1983) as

- iw ( t - x / c ) ¸
ÔÏ -iwroQe o
Ô - ikx.d /2|x| ikx.d /2|x|
p(x, t ) ª Ì ˝ (e -e ) (10)
ÓÔ 4p x ˛Ô

where k = w/co = 2p/l is the acoustic wavenumber, and l is the acoustic wave-
length. The first term in {} represents the sound field from one of the sources in
isolation, and the second term gives the effect of the difference in propagation
distance from each source to the observer. Combining the exponentials in the last
term gives -2i sin(kx . d/2|x|), which (if the jets are close together compared to
the acoustic wavelength so kd 1), can be approximated by -ikx . d/|x|. This
factor is less than kd, which is a small number, and so the sound field is signifi-
cantly reduced by combining the synthetic jets in this way. However, if the distance
between the jets is much larger than the acoustic wavelength then the factor of kd
is large and the sound field will depend on 2i sin(kx . d/2|x|). The resulting sound
field will be very directional with a maximum level which is twice the amplitude
of the sound level from one jet at the same distance. Therefore to reduce the sound
level by using synthetic jets in antiphase the distance between them must be very
much less than the acoustic wavelength at their excitation frequency.
Of course the desirable indirect aeroacoustic effects of synthetic jets may often
be the purpose of the control. One example that has reached full scale flight dem-
onstration is the active control of the wake shed by an avionics pod mounted
beneath the fuselage of an F-16 (Shaw et al., 2005, Maines et al., 2006, 2007).
Wake shedding from the pod causes lateral oscillations of the wake and excessive
acoustic loads on the downstream ventral fin that ultimately result in structural
failure. Six powerful synthetic jet actuators (Saddoughi, 2004) were installed just
upstream of the blunt base of the pod, which has a cylindrical body and hemi-
spherical nose. Wind tunnel tests were used to assist in selecting the operating
frequency and velocity of the synthetic jet that would suppress the wake motions
and thus the large unsteady pressures generated on the downstream ventral fin.
Pressure measurements made downstream of the pod in these tests clearly show
the tone noise generated by the synthetic jet actuators, but under full-scale condi-
tions this tone would be inaudible compared to the noise of the aircraft’s jet
engines. Flight testing showed reductions in pressure fluctuations on the fin by
65% at a Mach number of 0.5, demonstrating the effectiveness of this control.
Further development of this technology includes improvements to the actuators,
making them robust in adverse environmental conditions (including dust and
rain). Patenting of applications to wind turbines, aircraft engine tone noise control,
and operations in water and other liquids is underway.

IX. Summary
The objective of this chapter has been to show that there is a need to consider
the aeroacoustic implications of flow control devices, even if the primary objec-
tive of the control device is not to reduce noise. We have given a brief overview of
372 W. DEVENPORT AND S. GLEGG

the fundamental concepts of how sound is generated by flow, and discussed the
importance of the most common sources of aerodynamic sound. The coupling
between the unsteady flow and the radiated acoustic waves was discussed and it
was shown how the noise from flow disturbances in the presence of impermeable
structures was much more significant than the noise from the flow by itself. We
have illustrated these concepts with a review of the aeroacoustic implications of
some existing flow control devices, such as the higher harmonic control of heli-
copter rotors, rotor–stator interactions in duct fans, TE blowing, circulation control
using the Coanda effect, and synthetic jets. In some cases the flow control devices
are designed to reduce noise but in other cases, such as synthetic jets, the flow
control device introduces significant levels of unwanted sound, which may limit
their application to equipment for which noise is not an issue. In domestic equip-
ment, for example, noise could be a controlling design parameter and the use of
noisy flow control devices may not be acceptable. Understanding the aeroacous-
tics of flow control can therefore be a critical design consideration that needs to be
understood during the engineering development of a new device.
Chapter 12

Air-Breathing Propulsion Flowpath Applications

Daniel N. Miller*
Lockheed Martin Aeronautics Company, Fort Worth, Texas
and
Jeffrey D. Flamm†
NASA Langley Research Center, Hampton, Virginia

I. Tomorrow’s Propulsion Flowpath and the Need for Flow Control


Modern aircraft are characterized by a traditional wing/body/tail arrangement,
are all commanded by an onboard pilot, and are driven by the requirement for high
aerodynamic performance. As we look to the future, new aggressive requirements
for next-generation air vehicle systems are emerging which will undoubtedly both
accelerate and modify this design evolution. Examples of these future require-
ments include goals for an unprecedented level of range and loiter capability,
component integrability, and reliability, while accommodating stringent reduc-
tions in weight, volume, drag, noise, and fuel burn—all this in the context of a
tailless planform. Many of these requirements extend to both civilian and military
aircraft. Rising fuel costs, concern for the environment, and airport encroachment
are leading to greater emphasis on reductions in fuel burn, emissions, and noise in
the commercial sector. The NASA subsonic fixed wing program seeks to dramati-
cally improve noise, emissions, and aircraft performance by enabling changes in
engine cycles and airframe configurations as illustrated by the “N + 2” hybrid
wing/body configuration by 2020 (see Figs. 1 and 2). The corners of the N + 2
trade space include -52 dB cumulative noise reduction below stage 3 require-
ments, 80% reduction in LTO NOx emissions (relative to CAEP 2), and a 40%
reduction in fuel burn and field length (relative to a Boeing 777 with GE90

Copyright © 2008 by Lockheed Martin, Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*Technical Fellow, Air Vehicle Sciences and Systems. Associate Fellow AIAA.
†Aerospace Engineer, Configurations Aerodynamics Branch. Associate Fellow AIAA.

373
374 D. N. MILLER AND J. D. FLAMM

Fig. 1 The needs for tomorrow’s air vehicles pose design challenges that will require
advances in propulsion flowpath technology to meet them.

Fig. 2 NASA Fundamental Aeronautics: Subsonic Fixed Wing Program goals


(Alonso, 2007).
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 375

engines) (Alonso, 2007; NASA, 2007). These goals are aggressive and competing,
forcing the designer to trade requirements based on mission goals. Cambridge
University and Massachusetts Institute of Technology (MIT) have created a part-
nership called the silent aircraft initiative to explore the feasibility of designing a
commercial transport aircraft that would produce no perceived noise outside the
airport boundary (Crichton et al., 2007; De la Rosa Blanco et al., 2007; Hileman
et al., 2007a; Nickol, 2008; Plas et al., 2007). Future commercial transport designs
are expected to use many technologies including upper surface mounted embed-
ded engines with boundary-layer ingestion, variable area nozzles, and thrust vec-
toring to achieve these goals. These future air vehicle requirements represent a
daunting challenge for air vehicle system designers and technologists.
In this context, the aerodynamic design and integration of turbofan-engine
inlet and nozzle flowpath components will play a major role in defining the
configuration, capability, and cost of next-generation air vehicle systems. This
air vehicle evolutionary process and its associated propulsion flowpath aero
design challenges is illustrated in Fig. 1. To enable the necessary advance in air
vehicle design, tomorrow’s inlet and nozzle flowpath components must deliver
exceptional aerodynamic performance and functionality, while reducing weight,
size, drag, and fuel burn compared to state-of-the-art systems. Examples of
aerodynamic design challenges to meet these requirements include implementa-
tion of the following traits: a compact propulsion flowpath that both conforms
to the vehicle forebody surface and is embedded within the vehicle structure; a
simplified/lightweight thrust vectoring technique; concepts for reduced mechani-
cal complexity for articulation of moving inlet/nozzle surfaces; all this while
achieving high levels of inlet/nozzle aero performance. With these challenges in
view, the question has been posed—can a traditional view of aerodynamic
design and technology for the propulsion flowpath be embraced which meets
these challenges?
While a final answer for this question remains unanswered, designers and tech-
nologists in academia, government, and industry are investigating new techniques
and technologies that may enable implementation of these design features (e.g.
Allan et al., 2006; Brear et al., 2003; Miller et al., 1999, 2004). For example, air-
frame configurators are exploring the use of highly integrated (or embedded),
fixed-geometry inlet and nozzle flowpath designs to slash weight, volume, com-
plexity, and cost. Conventional aerodynamic design practices for inlets and nozzles
can preclude the use of an embedded inlet and nozzle system needed to meet these
future design goals. In this chapter, several examples of emerging modern flow
control technologies are presented in the context of a future inlet and nozzle flow-
path concept along with analyses, test results, and anticipated system benefits,
where available.

II. Applications to Tomorrow’s Propulsion Flowpath


The objective in this chapter is to briefly introduce some examples of future
inlet and nozzle design concepts and the vital role that modern flow control tech-
nology could play in their performance and function.
The propulsion flowpath of a modern, high-performance air vehicle can be
defined by a series of separate components (illustrated in Fig. 3). The jet engine
376 D. N. MILLER AND J. D. FLAMM

Fig. 3 Basic propulsion flowpath elements and their function.

is the heart of the air-breathing propulsion system, and operates by burning (often
hundreds of pounds per second of) air mixed with fuel, and turning that hot air
into thrust. The inlet pre-compresses and delivers that air to the jet engine,
whereas the nozzle is the component through which the hot air/exhaust products
exit the engine and are accelerated to produce thrust for the air vehicle. The
scope of this chapter will be confined to discussion of the non-turbomachinery
propulsion flowpath components, as turbomachinery applications are covered in
another chapter.
For the purposes of this chapter, the inlet opening region (or aperture) will be
distinguished from the inlet diffuser (or duct). The inlet aperture for a transonic/
supersonic-capable air vehicle is designed to capture efficiently the airflow needed
for the engine, provide some initial shock-based precompression, and finally to
divert or remove boundary layer that could be ingested into the inlet from the
forebody. To provide these functions, the modern inlet aperture contains a
boundary-layer splitter (or diverter) to divert the forebody boundary layer (see
Figs. 3 and 4). A series of variable geometry ramps may be employed to set up a
shock system intended to precompress the airflow. Furthermore, a large array of
boundary-layer bleed holes may be employed in the aperture region to stabilize
this inlet shock system from the adverse, unsteady effects of shock/boundary-
layer interaction. In total, the diverter, bleed, and variable shock-ramp features
provide for efficient transonic/supersonic inlet operation, but at a significant
weight, complexity, and cost penalty. The associated inlet duct is designed to dif-
fuse and deliver the airflow to the engine. To provide high efficiency, the modern
inlet duct is designed with minimal axial curvature and possesses a relatively
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 377

Fig. 4 Conventional inlet aperture/ramp bleed holes and boundary-layer diverter


illustrated on F-15.

low-aspect-ratio cross-sectional shape. To accommodate these features, the mod-


ern inlet duct is relatively long, generally expressed as inlet length/exit diameter
(L/D) on the order of 6, and is often mounted beneath or beside the forebody/
cockpit. Finally, the nozzle is designed to efficiently generate throttle and, in the
case of the modern high performance aircraft, vector thrust. A more detailed
description of modern propulsion flowpath system components may be found in
textbooks by Mattingly (2006), Mattingly et al. (2002), Oates (1989), and Seddon
and Goldsmith (1999).
A next-generation all-wing, air vehicle design concept with an embedded pro-
pulsion flowpath is illustrated in Fig. 5. Note that the inlet aperture is integrated
on the upper surface of the vehicle (referred to as top-mounted in this chapter, as
opposed to side- or bottom-mounted apertures in today’s air vehicles). Future
concepts and enabling technologies are being developed for a next-generation
inlet aperture, inlet duct, and nozzle system that will meet the aggressive needs
outlined previously. An advanced inlet aperture concept is envisioned that requires
no boundary-layer diverter, little or no boundary-layer bleed, and no variable
378 D. N. MILLER AND J. D. FLAMM

Fig. 5 Next-generation air vehicle concept with embedded propulsion flowpath.

ramp geometry. An embedded compact inlet duct is envisioned that is very short,
lightweight, structurally integrated into the airframe, and can accommodate high-
aspect ratio cross-sectional shapes and tight turns helping to integrate the duct
among bulkheads and subsystems. The embedded fixed-geometry nozzle should
be lightweight, structurally fixed, and still able to provide thrust vectoring (and/or
throttling) capability for supplemental control power on a tailless planform.

III. Application to the Inlet Aperture: Shock/Boundary Layer


Flow Control
As mentioned in the previous section, inlets that have supersonic capability
traditionally employ boundary-layer removal techniques. The techniques are
effective, but typically heavy, expensive, and they increase propulsion system
drag. Many aircraft use a boundary-layer diverter system such as the one shown in
Fig. 4. By design, forebody-induced boundary-layer or flow separation is diverted
away from the inlet. This system of boundary-layer removal is very effective. The
inlet will not experience the viscous losses of the ingested forebody boundary-
layer or any shock/boundary-layer interaction (SBLI) that might be present. The
inlet at supersonic speeds will primarily experience the total pressure loss through
the normal shock. However, the boundary-layer diverter will increase the drag,
weight, and cost of the vehicle. Another option that has been used is boundary-
layer suction. An example of a supersonic aircraft that uses both a boundary-layer
diverter and boundary-layer bleed (or suction) is shown in Fig. 4. Suction works
very well, and can bleed nearly all of the low energy flow in the boundary layer.
However, the boundary-layer bleed system possesses an inherent weight and
volume penalty. Also, most of the airflow bled out of the boundary layer is dumped
overboard, increasing propulsion system drag. Therefore, a technology is envi-
sioned that could eliminate the need for boundary-layer diverters and suction at
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 379

supersonic speeds, and could be adapted for future top-mounted, embedded inlets.
SBLI control is a field of study where a considerable amount of effort and research
has been performed. Only few of these efforts on SBLI control were focused on inlet
apertures (Wong and Hall, 1975). Some of the alternate bleeding methods that have
been used to control the interaction are tangential blowing (Schwendemann and
Sanders, 1982), conventionally sized vortex generators (VGs) (Mounts and Barber,
1992) and some more recent passive and active techniques including: slots
(Holden and Babinsky, 2003), low-profile, micro-VGs (Lin, 2002), and mesoflaps
(Hafenrichter et al., 2001).
Bruce and Babinsky (2007) investigated the use of passive, sub-boundary layer
VGs (micro-VGs) for SBLI control. Examples of these passive flow control actua-
tors are illustrated in Fig. 6 and are similar to those described by Lin (2002) in that
their height is between 10% and 50% of the boundary-layer height. High-speed
images from an experimental investigation (highlighted in Fig. 7) conducted at
Cambridge University indicated that the application of passive micro-VGs can
eliminate shock-induced separation (Ogaway and Babinsky, 2006).
This section will highlight an extension of these previous efforts using microjet
flow control on the forebody of an unmanned aerial vehicle (UAV) at Mach 1.4,
providing more efficient supersonic dash capability (Young et al., 2005). Microjet

Fig. 6 Example of vane micro-VGs (or micro-vanes) used for sub-boundary-layer


flow control. a) Courtesy of J. Lin, NASA LaRC; b) courtesy of H. Babinsky,
Cambridge University.
380 D. N. MILLER AND J. D. FLAMM

Fig. 7 Sample experimental results using ramp-style vane micro-VGs (or micro-
ramps) for shock/boundary-layer control. Courtesy of H. Babinsky, Cambridge
University.

flow control actuators, illustrated in Fig. 8, are consistent with the definition of
Lin (2002) where the jet diameter is 10–50% of the boundary-layer height. These
flight speeds cause a significant SBLI on this particular top-mounted inlet aperture
configuration, as shown in the computational fluid dynamics (CFD) analysis
illustrated in Fig. 9. Microjet flow control, considered an AFC technique, will be
evaluated to see if the addition of energy and vorticity inside the shock bifurcation
or lambda region can reduce the adverse effects of the SBLI. The flow control
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 381

Air Jet Micro-VGs (or Micro-Jets)

VGJ

Ue Pitch

Skew

Surface

Fig. 8 Illustration of air jet micro-VGs (or micro-jets) used for sub-boundary-layer
flow control. Courtesy of L. Cattafesta, University of Florida.

configuration will also be evaluated at off-design conditions to test the robustness


of the system.
The inlet/forebody shown in Fig. 9 was analyzed at a freestream Mach number
of 1.4 at 36,000 ft using a structured, Reynolds-averaged Navier–Stokes (RANS)
solver. Figure 10 shows the total pressure contours along the centerline of the

Fig. 9 Example application of microjet flow control to UAV inlet aperture/ramp


for shock/bounday-layer control.
382 D. N. MILLER AND J. D. FLAMM LIVE GRAPH
Click here to view

5.0% 5
Recovery
Efficiency
Normalized Total Pressure Change

4.0% 4

3.0% 3
% PT2 / PT0

Efficiency
2.0% 2

1.0% 1
∆Recovery
Efficiency =
% Microjet Mass Flow
0.0% 0
0.0% 0.5% 1.0% 1.5% 2.0% 2.5%
Injector Mass Flow

Fig. 10 Example CFD-based predictions using microjet flow control applied to UAV
inlet aperture/ramp for shock/boundary layer control.

forebody/inlet aperture region. The analysis shows that the classical normal shock
associated with an external compression inlet (see Seddon and Goldsmith, 1999).
One can also observe the rapid boundary-layer growth after passing through the
normal shock pressure gradient.
A microjet array flow control technique was defined parametrically using CFD
to guide the design process. Results showed that an array located within and across
the shock bifurcation region produced the best results (largest increase in total
pressure recovery for a given injected mass flow). The arrays were swept to align
with the normal shock around the 3-D inlet aperture. The microjets were both
inclined to the forebody surface and skewed relative to the center line. The effect
of microjet mass flow was examined by varying the total microjet array mass flow
from 1 to 2.25%.
The results shown in Fig. 10 show that the microjet mass flow has a significant
impact on inlet total pressure recovery. Total pressure recovery values cited were
computed using 160 data points at the inlet exit/engine face aerodynamic interface
plane (AIP). The effectiveness (% recovery/% microjet mass flow) of the flow
control peaks near a value of 2.5 at approximately 1.5% microjet mass flow. The
figure also shows that continuing to increase the microjet mass flow past 1.5%
results in a change in slope of the recovery and a decrease in the flow control
effectiveness. The inflow pressure of the microjet arrays simulated that in the
range of engine-fan-discharge pressure at the 1.5% design condition (Young et al.,
2005).
This type of flow control approach could be kept off during subsonic flight and
activated during supersonic flight conditions to maintain high performance and
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 383

stabilize the inlet shock system. Future prospects include a plasma-based flow
control actuation technique (Shin et al., 2006).

IV. Application to the Inlet Duct: Separation and


Vortex Flow Control
As has been suggested, the propulsion system is critically important in terms of
cost, weight, volume, performance, and overall configuration integration. In the
notional, tailless unmanned air vehicle (UAV) depicted in Fig. 5, the embedded
propulsion system—from inlet aperture to nozzle—is nearly as long as the aircraft
itself. The length of the airframe at the centerline is driven by the length of the
propulsion system. If the length of that system can be decreased, the entire air
vehicle can potentially be decreased in size. Since air vehicle costs scale propor-
tionately to weight, driving the size—and hence, weight—downward serves only
to decrease the overall cost (see Hamstra et al., 2000; MacMartin et al., 2001).
All the embedded propulsion system components contribute significantly to
propulsion system length (inlet, engine, nozzle). The length of the inlet system, in
particular, is critical not only to the air vehicle length, but also plays a major role
in determining the vehicle center of gravity. Inlet systems of current-generation
combat air vehicles are typically 10–20 ft long and can weigh on the order of
500–1200 lb. Two major figures-of-merit are used to measure inlet performance:
pressure recovery (the ratio of local and freestream total pressures), which gives
the overall efficiency of the system; and distortion, which gives the pressure non-
uniformity at the inlet/engine interface (see Fig. 11). High-pressure recovery is
desired to minimize fuel burn at a desired thrust level. Severe distortion can cause

Fig. 11 The basic functional performance metrics for an inlet; comparison of


conventional to embedded compact inlet duct.
384 D. N. MILLER AND J. D. FLAMM
LIVE GRAPH
Click here to view
1.0
Conventional Top CL Bottom CL

0.9 Duct
Surface Static Pressure Ratio

0.8

0.7
(Ps/ Pt0)

Side View LL
Embedded
0.6 Compact Duct Comparison X
Embedded
0.5 Compact Duct

Conventional
Second Turn
0.4 Duct

First Turn
0.3
0.0 0.2 0.4 0.6 0.8 1.0
Axial Location (X / L)

Fig. 12 Axial pressure gradient comparison between conventional and embedded


compact duct.

stall and possibly flameout of the jet engine. A comparison of CFD-based wall
pressure distribution is shown in Fig. 12 for the embedded compact inlet duct
compared to a conventional inlet duct. Even with optimization, the embedded
compact duct exhibits a much larger magnitude and rate of streamwise wall
pressure rise than that of the conventional inlet duct, indicating the potential for
massive flow separation and high losses. The fluid mechanics of an embedded
compact (serpentine) inlet duct is illustrated in Fig. 13. An isometric view of a
CFD solution is shown at the top (13a), while a plan view of experimental flow
visualization results on a similar duct are shown below (13b). This serpentine duct
is characterized by severe wall curvature that induces strong secondary flows,
resulting in a coalescence of the boundary-layer on the inner wall, and subsequent
vortex lift-off separation. Such phenomena produce unacceptable levels of pres-
sure loss, flow distortion, and turbulence.
To meet vehicle weight and cost reduction goals, future inlet ducts must be
lighter, more compact, and must accommodate ever-increasing integrability
design requirements between the air vehicle and engine (see Fig. 1). However,
future design concepts for the embedded compact inlet duct length, degree of
curvature, and cross-sectional shape are limited by considerations of pressure loss
and flow non-uniformity (distortion). As future systems evolve toward more com-
pact designs with exotic, embedded flowpath shaping, these limitations will in
turn limit the design space for the vehicle itself. A need thus exists for new tech-
nologies that can overcome these inlet design limitations.
Researchers within the government, academia, and industry are turning toward
the emerging technology of AFC to address these problems. Historically, the most
common method of flow control in inlet ducts has been the inclusion of vane or
air-jet type VGs to “locally” control the adverse effects of separation. Vorticity
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 385

Fig. 13 CFD solution and experimental flow visualization showing development of


axial-curvature induced vortices in an S-duct inlet (Top: courtesy of John Sullivan,
Purdue University).

generation from the VGs is used to locally mix low- and high-momentum regions
in the flow, effectively spreading out the lower momentum fluid to suppress flow
separation from the wall (see Tindell, 1987). However, application of this flow
control method to advanced serpentine inlet ducts does not necessarily achieve
significant reduction of engine face distortion. Furthermore, the local use of VGs
only allows separation to be controlled at one flow condition (usually the cruise
condition), with all other conditions rendered “off-design” (Anderson and Gibb,
1998). More recent studies have similarly addressed control of separation in a 2-D
duct using pulsed or synthetic jets (Amitay et al., 2000). Although reattachment of
the separated flow is obtained, such studies are mainly of academic interest
because they were conducted on simplistic flowpath geometry at very low Re
conditions, demonstrating no direct benefit to inlet system-level metrics.
Owens et al. (2006) performed an experimental investigation on the effect of
active and passive flow control (PFC) on inlet distortion for an S-inlet with
boundary-layer ingestion at a Mach number of 0.85. This study documented
results of inlet duct flow control at conditions representative of flight Mach
numbers over a range of inlet operating conditions. The authors demonstrated a
386 D. N. MILLER AND J. D. FLAMM

reduction in SAE circumferential distortion (see SAE Document ARP1420 Rev B,


Gas Turbine Engine Inlet Flow Distortion Guidelines, March 2002) from 0.055 to
as low as 0.015 with control jet blowing 2.5% of inlet mass flow. At the inlet
design point the VGs reduced the circumferential distortion from 0.55 to 0.10
(Fig. 14).
Modern flow control is characterized by the exploitation of very small-scale (or
microscale) perturbations to the flowfield, usually near the wall or other flowfield
boundary to control large-scale aerodynamic “global” flow phenomena. In this
section, results are given of an effort to design and verify an embedded compact
inlet duct using a passive micro-VG flow control technique to provide high perfor-
mance levels. This passive vortex control technique seeks to suppress the growth of
larger-scale vortices generated in a serpentine inlet duct with the goal of simultane-
ously improving the inlet system level performance metrics of total pressure recov-
ery, spatial distortion, and RMS turbulence (see Fig. 15). In this approach, separation
control is a secondary benefit, not a design requirement. The embedded compact
inlet for this study was a 4:1 aspect ratio ultra-compact (L/D = 2.5) serpentine duct
(see Fig. 16). Two sets of passive micro-VG flow control arrays were designed with
the intent of establishing high performance levels to the baseline duct. The flow
control design included two arrays of 36 co-rotating microvane VGs. Optimization
of the microvane array was accomplished using a design of experiments (DOE)
methodology to guide the selection of parameters used in multiple CFD flow solu-
tions, as illustrated in Fig. 17 (Hamstra et al., 2000). An example of DOE-derived
design sensitivity is shown in Fig. 18.
A verification test conducted in the NASA Glenn W1B test facility indicated
low-pressure recovery and high distortion for the baseline duct without flow

Fig. 14 Boundary-layer ingesting flow control experiment (Owens et al., 2006).


AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 387

Fig. 15 Microvane-based vortex flow control technique illustrated with CFD solutions.

control. Experimental engine face total pressure patterns are shown in Fig. 19.
Results across a range of inlet Mach Numbers (representing a wide range of
throttle settings) are shown in Fig. 20. With microvane flow control, at a throat
Mach number of 0.60, pressure recovery was increased 5%, and both spatial dis-
tortion and turbulence were decreased approximately 50%. A comparison of CFD

Fig. 16 Microvane-based vortex flow control technique design variables.


388 D. N. MILLER AND J. D. FLAMM

Fig. 17 Process used to develop inlet flow control design concept couples CFD, DOE
methods, and test evaluation.

prediction and experimental data for the microvane configuration is shown in


Fig. 21, with both engine face pattern and longitudinal static pressure distribution
illustrated. In this case, the microvane effectors have controlled the flow separa-
tion, and the test data and CFD analysis compare very well. Taken together, the
substantial performance improvement and favorable test-to-CFD comparison are

Fig. 18 Example DOE-based sensitivity (generated from CFD analyses) used to


design inlet flow control concept for reduced distortion.
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 389

Fig. 19 Experimental results showing inlet duct exit total pressure profiles for
baseline and case with microvane-based flow control.

LIVE GRAPH
Click here to view
0
Baseline
Normalized Total Pressure Change

Micro-Vanes
–2%

–4%
% PT2/PT0

–6%

–8%

–10%
0.40 0.45 0.50 0.55 0.60 0.65 0.70
Throat Mach Number

Fig. 20 Example performance characteristics from embedded compact inlet duct


testing both without flow control (baseline) and with micro-vane array flow control.
390 D. N. MILLER AND J. D. FLAMM LIVE GRAPH
Click here to view

Fig. 21 Comparison of CFD predictions vs test results for axial pressure distribution.

interpreted as verification of the microvane effector design methodology. Future


prospects for embedded compact inlet ducts, summarized in Fig. 22, include both
pulsed microjet injection (MacMartin et al., 2001) and synthetic, zero-net-mass-
flux (ZNMF) jet actuated flow control (Amitay et al., 2000).

Fig. 22 Future prospects for inlet duct flow control a) Courtesy of J. Paduano (MIT);
b) courtesy of A. Glezer (GT).
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 391

V. Application to the Nozzle: Fluidic Thrust Vectoring Flow Control


The weight and cost of tactical aircraft exhaust systems has increased at an
alarming rate with the incorporation of features for afterburning, thrust vectoring,
and advanced shaping. There is also interest in incorporating thrust vectoring into
future commercial hybrid wing/body aircraft designs such as the SAX-40. The
SAX-40 requires thrust vectoring for rotation at take-off and for pitch trim in lieu
of elevons during climb-out and at cruise, requiring as much as 10.5° of pitch
vectoring (Crichton et al., 2007; Hileman et al., 2007b).
Historically, afterburning and vectoring have required variation of the nozzle
geometry (illustrated in Fig. 23). A typical turbofan engine’s throat area must
increase in size when afterburning. Vectoring has required deflection of nozzle
divergent flaps, if not rotation of the entire nozzle assembly. Aperture shaping for
afterbody integration further imposes the use of less structurally efficient 2-D,
rather than axisymmetric, nozzles. These capabilities require greater mechanical
complexity in the system which are significantly (~50%) heavier than the less
capable nozzles of current tactical aircraft (Miller et al., 1999). Simply stated,
there is a technical need to obtain jet control (thrust vectoring and afterburning)
within the confines of a mechanically simple nozzle that retains as much fixed
structure as possible. This future nozzle concept is referred to as an embedded
fixed-geometry nozzle. Such simplicity reduces parts count, weight, and cost (see
Fig. 24). The objective of this simplified concept is to combine the performance
capabilities of a modern high-performance aircraft nozzle with the integration
benefits of a fully fixed nozzle by controlling the aerodynamic flow of the jet
using fluidic techniques. Fluidic thrust vectoring and area control is accomplished
by manipulating or controlling the primary exhaust flow with the use of a second-
ary air source, typically bleed-air from the engine compressor or fan.
Foundational work employing fluidic injection to control nozzle effective throat
area was conducted in the late 1950s and early 1960s. Historic literature addressed

Fig. 23 Basic functional performance metrics for a nozzle.


392 D. N. MILLER AND J. D. FLAMM

Fig. 24 A key advanced nozzle design: integration issues.

the basic effects of internal nozzle convergence, flow conditions, injector angle,
and injector location on discharge coefficient. However, injector mass flow rates
were considered excessive for application to afterburning conventional turbofan
engines. A review of this work is contained in Miller and Catt (1995). More recent
work on a fully fixed, fluidically controlled nozzle reduced fluidic injection flow
rate beyond previous efforts. Peak throat-area control effectiveness was obtained
with a slot injector integrated near the geometric throat and angled to be highly
opposed to the primary nozzle stream. In another effort, a fixed-aperture nozzle
concept was identified that varied effective throat area by injecting secondary air
at the nozzle throat and routing nozzle core air around the geometric throat to be
“re-injected” into the core stream (Miller et al., 1997).
An overview of advanced fluidic thrust vectoring concepts is found in an inlet
and nozzle technology paper by Gridley and Walker (1996) and a fluidic thrust
vectoring summary paper by Deere (2003).
Three primary mechanisms of fluidic thrust vectoring have been studied over
the last 15 years: shock-vector control, counterflow, and virtual aerodynamic
surface shaping. The later technique, aerodynamic surface shaping, can be further
divided into two methods: throat skewing and separation control. These techniques
can be used to vector the exhaust flow in the pitch or yaw direction (see summary
in Fig. 25a and b).

A. Fluidic Counterflow Thrust Vectoring


The counterflow thrust vectoring technique was reported by Strykowski and
Krothapalli (1993) and Strykowski et al. (1993, 1997). This concept used a
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 393

Fig. 25. Overview of fluidic nozzle flow control techniques.

suction source applied adjacent to the primary nozzle. The primary nozzle is
shrouded by a suction collar (nozzle flap, see Fig. 25a), and the presence of the
suction collar creates a highly overexpanded nozzle. The applied suction increases
mixing between the secondary and primary flows creating a countercurrent shear
layer between the two flows. The presence of the collar inhibits mass entrainment
and the flow near the collar accelerates, causing a drop in pressure on the collar.
If the vacuum is applied asymmetrically to one side of the nozzle or the other, the
jet will vector toward the low-pressure region (Flamm, 1996, 1998). The
countercurrent shear layers exhibit higher turbulence and mixing levels than
394 D. N. MILLER AND J. D. FLAMM

coflowing shear layers (Strykowski et al., 1997). The enhanced mixing character-
istics of the counterflow nozzle may have the added benefits of reducing jet noise,
jet temperature, and emissions from the nozzle (Strykowski et al., 1993). Thrust
vector angles of 16 deg were reported at relatively small suction mass flow rates
(Strykowski et al., 1997). Larger-scale experimental studies by Flamm (1996,
1998) and computational studies by Hunter and Deere (1999) confirmed the
nozzle thrust vectoring performance, but found that the concept exhibited hyster-
etic jet attachment to the suction collar (nozzle flap). These experiments also
reported a range of thrust efficiencies of 0.92–0.97. The counterflow concept
does not lend itself to direct comparison with other fluidic thrust vectoring tech-
niques through the traditional means of reporting vectoring efficiency (degrees
vectoring/percent secondary injection flow rate). However, the secondary suction
flow rates are very small and typically less than 1% of the primary flow (Flamm,
1998). Future challenges for this concept must address a viable suction source,
as well as thrust loss, weight penalty due to the large suction collar, potential
vectoring instability associated with a highly overexpanded nozzle, and hyster-
etic jet attachment.

B. Fluidic Shock Vector Thrust Vectoring


Shock vector control (SVC) methods rely on the formation of a strong oblique
shock in the nozzle divergent section to achieve flow turning. The oblique shock
is generated by introducing an injected flow asymmetrically into the supersonic
primary nozzle flow (Fig. 25a). This approach achieves thrust vector control
without varying nozzle throat area. This method can produce large thrust-vector
angles, but at the expense of thrust efficiency as losses occur. Although this
method is effective in vectoring the nozzle stream, SVC is typically character-
ized by large thrust losses caused by the primary flow passing through the oblique
shock. The oblique shock may also impinge on the opposing nozzle wall at
certain operating conditions, which results in thrust vectoring and thrust perfor-
mance penalties (Deere, 2003). The SVC method has been shown to produce
thrust vectoring efficiencies in the range of 0.9–4 deg per percent of primary
injected mass flow with system thrust efficiencies ranging from 0.86 to 0.94
(Deere, 2003). For further discussion of SVC concepts, see Abeyounis and
Bennett (1997); Anderson et al. (1997); Chiarelli et al. (1993); Deer and Wing
(1998); Deere (2000); Federspiel et al. (1995); Giuliano and Wing (1997); Waithe
(2001); Wing (1994); Wing et al. (1997).

C. Virtual Aero Surface Shaping via Throat Skewing


Virtual aero surface shaping for the embedded fixed-geometry nozzle is accom-
plished by injecting secondary air in the nozzles divergent section or at the throat
(Figs. 25b and 26) with the goal of creating “virtual aerodynamic surfaces.” The
virtual surfaces provide the same functionality as today’s variable-geometry
mechanical flaps (Miller et al., 1999). The nozzle’s aerodynamic throat is shifted
or skewed by the virtual aerodynamic surfaces. By injecting from both nozzle
flaps symmetrically about a given axial location, the minimum flow area of the jet
exhaust can be changed, thereby allowing for changes in engine power setting.
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 395

Fig. 26 Overview of fluidic virtual aero-surface nozzle flow control thrust vectoring
concept.

Injecting asymmetrically (either at different axial locations or from only one flap),
results in an asymmetric pressure distribution on the two nozzle flaps, thereby
creating vectored thrust.
A DOE-based process (similar to the embedded inlet) was used to develop the
nozzle system flow control for the above virtual aero surface throat skewing
concept. DOE methods are used to search through the design space, and CFD is
used to evaluate the thrust coefficient and vectoring capability of each design
element. Subscale tests are periodically conducted to validate the design process
(Miller et al., 1999). An example design sensitivity for vectoring efficiency versus
location of the injector on the nozzle flap resulting from the DOE process is shown
in Fig. 27. Moving the secondary flow injector aft on the flap toward the nozzle
exit has a strong effect to increase thrust vectoring efficiency.
Since thrust vectoring is used as part of the flight control system, bandwidth or
response rate is of concern to ensure that forces can be applied at the proper time
scale. A brief study was done to predict the transient dynamic thrust vectoring char-
acteristics to ensure that the response rate of the fluidic virtual aero-surface thrust
vectoring technique would be fast enough for aircraft control. Based on an unpub-
lished internal Lockheed Martin investigation, unsteady CFD results predicted a
rate of thrust vectoring angle change well in excess of 100 deg/s using a step input
function for pressure at the injector location. This ideal level of bandwidth is more
than an order of magnitude higher than most aircraft control systems require.
Therefore, a significant margin is available for incorporating injector valves and
control hardware, which will certainly reduce the effective bandwidth.
The experimental demonstration of a multi-axis thrust vectoring (MATV) flu-
idic virtual aero-surface thrust vectoring via throat skewing is highlighted in Figs.
28 and 29 for an embedded fixed-geometry nozzle (Yagle et al., 2000, 2001). This
nozzle is representative of the type of nozzle installed in the next-generation
vehicle concept shown in Fig. 5. A demonstration of MATV was conducted on a
20%-scale embedded fixed-geometry nozzle test article across a range of nozzle
396 D. N. MILLER AND J. D. FLAMM

Fig. 27 Example of DOE-based sensitivity (generated from CFD analyses) used to


design fluidic virtual aero-surface T/V nozzle flow control concept.

pressure ratios (NPR), injector flow rates, and flow distributions. The stereolithog-
raphy nozzle is pictured in Fig. 28. Data for this nozzle are shown in Fig. 29. At
this high power setting, NPR = 5.5 (representative high-subsonic cruise), yaw
vector angles in excess of 7 deg were obtained while still maintaining a high thrust
efficiency (Cfg) in excess of 0.94. Higher yaw vector angles, in excess of 13 deg,
were achieved at the low power setting, NPR = 2.0 (representative of a nozzle
pressure ratio required for take-off), with a 2.5% penalty in thrust efficiency com-
pared to the high power setting. Similar results were obtained for pitch vectoring
measurement, but are not shown here. A detailed discussion can be found in the
paper by Yagle et al. (2001). Values of thrust-coefficient for the virtual aero-surface
T/V via throat skewing generally exceeded published measurements of shock-
based, vectoring methods. In terms of vectoring effectiveness (ratio of vector
angle to percent injected flow), fluidic throat skewing was found to be comparable
to shock-based vectoring methods.
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 397

Fig. 28 3-D fluidic thrust vectoring nozzle geometry.

Yaw Vector Angle Gross Thrust Coefficient


14 Low Power 1.00

12 High Power 0.98


0.96
10
0.94
8
Cfg 0.92
δy 6
0.90
4
0.88
2
0.86
0
0.84
–2 0.00 0.01 0.02 0.03 0.04 0.05
0.00 0.01 0.02 0.03 0.04 0.05
ω√τ ω√τ

LIVE GRAPH Fig. 29 Experimental results of fluidic nozzle T/V. LIVE GRAPH
Click here to view Click here to view

D. Virtual Aero Surface Shaping via Separation Control


An alternate virtual aero-surface shaping technique, which uses separation
control as means to vector thrust, has been researched at NASA Langley Research
Center (Deere et al., 2003, 2005, 2007; Flamm et al., 2005, 2006, 2007). This
technique is referred to as the dual throat nozzle (DTN) concept because of the
two minimum areas that form a convergent-divergent-convergent nozzle shape
(see Figs. 30 and 31). The DTN design enhances the throat shifting method of
thrust vectoring using fluidic injection at the upstream minimum area to control
separation and maximize pressure differentials in the cavity. In the non-vectoring
mode, no fluidic injection occurs and the sonic line or “throat” of the nozzle lies
at the nozzle exit minimum area. In the thrust-vectoring mode, asymmetric fluidic
injection is introduced at the upstream minimum area, causing a skewed sonic line
upstream of the exit and an increased pressure differential along the cavity walls.
398 D. N. MILLER AND J. D. FLAMM

Separation Control Dual Throat Nozzle


Sonic line

Characteristics
• New nozzle shape
No Injection, Un-vectored • Asymmetric injection
• Throat shifting plus enhanced
pressure differential
Sonic line • Highest published thrust vectoring
efficiency with high thrust efficiency

Vectored

Fig. 30 Overview of fluidic thrust vectoring concept using separation control.

Axisymmetric and 2-D DTN concepts have been extensively studied both
computationally and experimentally. CFD was used to guide a series of parametric
studies exploring the DTN geometric design variables including cavity shape,
upstream and downstream throat heights, secondary flow injection angle, conver-
gent and divergent ramp angles, cavity length, and secondary injector shape.

Fig. 31 Dual Throat Nozzle (DTN) design variables and installation photos.
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 399

In the case of the axisymmetric nozzles, circumferential extent of secondary


injection and expansion ratio were also studied. The computational studies were
then validated through experimental testing (see Figs. 31 and 32).
The DTN concept is capable of producing thrust vector angles over 18 deg at
NPR = 4 while maintaining high thrust efficiencies (Cfg, sys) between 0.92 and 0.96
(Flamm et al., 2006). At low to moderate injection rates (3% and below), the thrust
vectoring efficiency (h) of the DTN concept ranges from 4 to over 12; this is the
highest vectoring efficiency of any published fluidic thrust vectoring concept
(Deere et al., 2007; Flamm et al., 2006, 2007). A concern with the DTN concept
is the large variation in nozzle discharge coefficient with secondary injection ratio.
As seen in Fig. 32, the nozzle discharge coefficient (Cd,prim) varies over 10% with
increasing secondary injection rate (and consequently thrust vector angle). Throat
area control will likely be required to control discharge coefficient variation with
the DTN concept. This is possible through the use of differential blowing at the
throat (blowing asymmetrically from both the upper and lower surfaces of the
throat), but may impact on thrust vectoring efficiency due to the increased use of
secondary air. Studies are ongoing at NASA LaRC to investigate the application
of throat area control to the DTN concept.
Figure 33 presents a comparison of experimental data for various fluidic thrust
vectoring concepts (Flamm et al., 2007). The virtual aeroshaping (throat skewing
and separation control) concepts generally exhibit better thrust efficiencies
than shock vector concepts except at high NPR (greater than 8) where the virtual
aero-shaping concepts are operating well off design. The SVC concept relies on
oblique shocks in the nozzle to turn the flow, thus incurring a performance loss
across the shock.

LIVE GRAPH
Click here to view

Fig. 32 DTN effect of cavity convergent ramp angle, q2 (Flamm et al., 2006).
400 D. N. MILLER AND J. D. FLAMM

LIVE GRAPH
Click here to view

Fig. 33 Comparison of fluidic thrust vectoring techniques (Flamm et al., 2007).

E. Pulsed Actuators and Reduced Injector Flow Rate


Fluidic thrust vectoring methods such as virtual aero-shaping and SVC rely on
a secondary flow to block the primary flow. By increasing the penetration of the
secondary flow into the primary jet crossflow there is a corresponding increase in
blockage and in turn thrust vectoring. All of these fluidic vectoring methods rely
on a source of injected air, typically bled from the engine. Since engine bleed air
reduces the performance of the aircraft, there is a desire to minimize the amount
of bleed air required for thrust vectoring. Research efforts have typically focused
on optimizing nozzle geometry to enhance blockage and increase vector control
for a given amount of injected mass flow. An alternate approach is to modify the
injection technique to increase the apparent blockage for a given injected mass
flow. There are three primary areas of investigation for enhancing injector penetra-
tion/apparent flow rate: the use of periodic excitation actuators, pulsed injection,
and pulsed ejection (see Fig. 34).
Pack et al. (1999, 2001a) demonstrated the use of periodic excitation using a
Piezoelectric actuator applied to a divergent nozzle at low speeds. Deflection
angles of the order of 8 deg were demonstrated using the electric actuator.
There has been relatively little published research on pulsed injection in high-
speed compressible flows. Miller et al. (2001) computationally studied the use of
pulsed injection to improve jet penetration in a convergent duct followed by a
constant area section. The flow was near Mach 1 at the throat of the convergent
duct at NPR 2. The pulsed injection was introduced in the constant area duct just
downstream of the convergent. The study found that increasing pulse frequency
and pulsing 45 deg upstream into the primary jet crossflow produced improved
blockage and penetration compared to steady blowing.
Deere et al. (2005) investigated the application of pulsed injection in a super-
sonic nozzle for thrust vectoring. The computational study looked at the effect of
frequency, amplitude, and duty cycle on secondary jet penetration and thrust
AIR-BREATHING PROPULSION FLOWPATH APPLICATIONS 401

Fig. 34 Key research areas to reduce injector flow rate for fluidic T/V nozzles.

vectoring in the DTN nozzle. The study showed that pulsing the secondary injec-
tion jet increased the penetration and blockage into the primary jet crossflow at
the peak of the pressure pulse. However, the penetration was not maintained
through the pulsing period as the injection pressure reduced. Consequently the
time-averaged thrust vector angle was not improved vs the steady-state case for
this application. Barruzzini et al. (2007) reported similar findings when applying
pulsed injection to throttle a supersonic rocket nozzle. Similar levels of flow con-
trol performance were obtained between steady and pulsed injection when time-
averaged mass, momentum, and energy fluxes were made equivalent.
An alternative to pulsed injection is pulsed ejection. The pulsed ejection tech-
nique proposed by Yagle et al. (2002) uses a pulsed high-pressure primary stream
to increase the entrainment of a co-annular secondary flow (see Fig. 34). This
device is used to increase the mass flow and hence blockage of an injection actua-
tor without increasing the requirement for high-pressure bleed flow from the
engine. The study found a 75% increase in pumping effectiveness for the ejector
actuator vs steady-state injection.

VI. Summary
The next generation of commercial and military aircraft are evolving to highly
integrated hybrid wing/body designs. This evolution is driven by requirements to
dramatically reduce fuel burn, noise, and emissions while increasing range and
loiter capability. These highly integrated designs necessitate highly integrated
propulsion systems with top-mounted boundary-layer ingesting offset-inlets as
well as thrust vectoring and nozzle area control. The application of flow control
402 D. N. MILLER AND J. D. FLAMM

concepts to the propulsion system will play a critical role in solving the challenge
of inlet distortion and integrating lightweight fixed nozzles to these future aircraft.
An overview of modern research into these flow control techniques for propulsion
flowpath integration has been included in this chapter. More research and devel-
opment is needed to render these techniques realizable for next-generation air
vehicle platforms.
Chapter 13

Flow Control for Rotorcraft Applications

Ahmed A. Hassan*
Sigma Technologies, Mesa, Arizona
Michael A. McVeigh†
Boeing Company, Philadelphia, Pennsylvania
and
Israel Wygnanski‡
University of Arizona, Tucson, Arizona

I. Background
Future rotorcraft and tiltrotor aircraft have to be faster, quieter, and aerodynami-
cally efficient. Additionally, for military aircraft, they have to be more agile and
less detectable. These requirements are dictated by the ever-increasing demand
for extra range, additional payload, increased passenger comfort and, for certain
applications, increased lethality and survivability To accomplish these goals,
order-of-magnitude improvements in rotor, fuselage, empennage, and engine
aerodynamics must first be demonstrated. With researchers recognizing that pas-
sive flow control (PFC) means are limited in meeting these requirements, emphasis
has been placed on the identification and validation of novel active flow/noise
control strategies and, more importantly, the development of suitable actuators
(flow effectors).
For rotorcraft, active flow control (AFC) has often utilized “steady” blowing
through slots in the tail boom or blade. This strategy has been utilized on the
McDonnell Douglas NOTAR™ anti-torque system seen on the MD520N series of
helicopters (Sampatacos et al., 1983). In the presence of the rotor downwash,
steady tangential blowing through two spanwise slots on a circular cross-section

Copyright © 2008 by the authors. Published by the American Institute of Aeronautics and
Astronautics, Inc., with permission.
*President. Associate Fellow AIAA.
†Senior Technical Fellow. Senior Member AIAA.
‡Professor, Aerospace and Mechanical Engineering. Fellow AIAA.

403
404 A. A. HASSAN ET AL.

Fig. 1 Typical helicopter flowfield environment (courtesy of Prof. J. G. Leishman,


The University of Maryland and Cambridge University Press).

tail boom is used to produce an aerodynamic side force due to the “Coanda effect.”
This force is used for directional control in lieu of the more commonly used, and
often dangerous, tail rotor.
In general, the flowfield of a rotary wing aircraft is highly unsteady and 3-D.
Figure 1 is a sketch illustrating a typical helicopter flowfield environment.
In low-speed descent flight, close encounters take place between one rotor blade
and the trailing tip vortex shed by the preceding blade. Sometimes the interaction
may take place between the blade and its own tip vortex. These encounters, more
dominant on the advancing rotor blade side than on the retreating side, are com-
monly referred to as blade–vortex interactions (BVI). Resulting from these interac-
tions are significant impulsive changes in the rotor blade aerodynamic loads which
lead to an increase in vibration and noise levels—both, of course, having adverse
effects on passenger comfort and community acceptance near heliports.
The use of tip air mass injection (TAMI) has also been investigated as a viable
method to alleviate BVI noise (Pegg et al., 1975). In this approach, a jet of air,
preferably aligned with the axis of the tip vortex, is introduced at the tip of the blade.
The primary objective is to modify the structure of the tip vortex and/or introduce a
flow instability that enhances its dissipation, thus reducing the intensity of the inter-
actions. Obviously, the success of this strategy is relies heavily on the ability to align
the tip air jet with the axis of the tip vortex, a task which demands that one knows a
priori the relative position of the tip vortex with respect to the blade.
The use of higher harmonic control (HHC) of blade pitch (Brooks et al., 1989)
is a different strategy to actively control the BVI noise and vibration levels of a
helicopter rotor. Here, blade root harmonic pitch oscillations are superimposed on
the primary rotor pitch motion to alter the local aerodynamics of the blade with
the net result of increasing the blade–vortex separation distance, and thus the
intensity of the interactions.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 405

In high-speed forward flight, on the advancing side of the rotor disk, the flow is
dominated by compressibility effects which give rise to the formation of shock
waves with their attendant wave drag, shock noise, and shock boundary-layer
interactions (SBLIs). On the retreating side of the rotor disk, the flow is dominated
by viscous effects where the phenomena of dynamic stall and the associated large
excursions in drag and pitching moments are known to limit the maximum attain-
able forward flight speed of the vehicle.
On the other hand, tiltrotor aircraft such as the Bell XV-15 and the Bell/Boeing
V-22 Osprey are revolutionary twin-rotor vehicles that combine both the low-
speed benefits of conventional helicopters and the high-speed cruise benefits of a
fixed-wing turbo-prop aircraft. In hover, the wakes from the two rotor systems
impinge on the upper surface of the wing resulting in what is commonly referred
to as a download force. This force, also known as vertical drag, adversely impacts
the aerodynamic performance of the vehicle and, specifically, its useful payload.
For the XV-15 aircraft, this penalty described by Maisel et al. (1986) can be as
high as 15% of the gross weight if the wing flaps are not deflected, and as low as
5% of the gross weight with the inboard and outboard flaps deflected downwards
to reduce the projected wing area.
In this chapter, examples are presented where the benefits from three control
mechanisms are demonstrated through the use of numerical simulations and
results from wind tunnel tests. It is shown that while a simple formulation such as
that based on the full potential equation is sufficient to provide the global effects
of the control, more comprehensive formulations such as those based on the more
comprehensive Euler and Navier–Stokes equations are essential for capturing the
local details of the inherently complex flow interaction phenomena. For brevity,
validation of the numerical tools is only provided in the cited references.

II. Applications of Active Flow Control


In this section we demonstrate the benefits of using controlled injection of
vorticity on the resulting global/local aerodynamic characteristics of an airfoil/
blade/wing. At this juncture, it is instructive to refer to the vorticity flux equation
given by Reynolds and Carr (1985):

∂w Ê ∂U S ˆ 1 Ê ∂p ˆ
-n a + - (V w )
∂n ÁË ∂t ˜¯ r ÁË ∂s ˜¯
vorticity surface pressure surface
flux acceleration gradient injection

In the preceding equation, w is the vorticity, n is the kinematic viscosity, n is the


direction normal to the aerodynamic surface (e.g., blade), US is the local velocity
of the surface, r is the fluid density, p is the local pressure, s is the streamwise
distance measured along the surface, and V is the normal surface transpiration
velocity (positive or negative to represent the application of blowing or suction
respectively).
The first term on the right implies that the created flux of vorticity is propor-
tional to the acceleration of the aerodynamic surface. This acceleration, for
406 A. A. HASSAN ET AL.

example, can be a result of an unsteady pitch-type motion, an oscillatory plunge-


type motion, or an oscillatory streamwise lead-lag motion.
The second term represents the streamwise pressure gradient over the aerody-
namic surface. The most influential parameter having a major impact on the rela-
tive magnitude of this term is the surface geometry (i.e., surface slope and
curvature). The impact of this term is examined in section B where an unsteady
fluidic bump is created to reduce the strength of a shock wave.
The third term represents the influx or efflux of steady and/or unsteady vorticity
through the use of blowing or suction, respectively. The impact of this term is
examined in sections A2–4 and sections B–D.
Note that while each of the terms on the right exemplifies, implicitly, the under-
lying physical mechanism inherent in the control, these terms also suggest the
broad control strategy to be adopted.

A. Helicopter Blade-Vortex Interactions (BVI)


Under certain flight conditions, and in particular during low-speed descent (see
Fig. 2), BVI take place resulting in significant impulsive changes in the rotor
blade aerodynamic loads which consequently lead to an increase in vibration and
noise levels (Boxwell et al., 1983).
The direct relation between the strength of helicopter BVI, and hence BVI noise
levels, and the temporal pressure gradients near the leading edge (LE) of a blade
have been widely recognized by researchers (Brooks, 1993; Caradonna et al., 1988;
Carlin et al., 1989; Dawson et al., 1995). In general, large gradients are indicative of
strong interaction(s) which typically result from the presence of relatively strong,
near-parallel vortex wake segments close to the surface of the blade. Quite often,
however, it is the differential pressures near the LE of the blade, or their temporal
gradients, that are used as a measure of the intensity of BVI. This latter quantity also

Fig. 2 Sketch and full potential results illustrating the interaction of a two-bladed
rotor with its tip vortex wake (OLS model rotor: Psi = 70 deg; Mtip = 0.666;
Rbar = 0.55; Mu = 0.163; as = -2 deg (Hassan, 1989).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 407

represents the impulsive changes in lift—thereby providing insight into the resulting
blade vibration levels during the blade–vortex encounters.
Motivated by the need to meet FAA certification requirements for allowable
BVI noise levels and to increase community acceptance near heliports, research-
ers attempted to alleviate the impulsive response of the rotor blade to the vortex
passage by disrupting the natural mechanisms that are responsible for promoting
these encounters. With passive control (Brooks, 1993; Carlin et al., 1989; Pegg
et al., 1975) only able to achieve incremental reductions (1–2 dB) in BVI noise
levels, researchers focused on exploring novel active BVI noise reduction tech-
niques (Brooks et al., 1989; Dawson et al., 1995). Fundamentally, whereas these
techniques are commonly perceived as “active noise control,” the noise reduction
benefits are a direct outcome of equally successful AFC strategies, since the
resulting acoustics of a blade are merely a byproduct of its aerodynamic environ-
ment (Dawson et al., 1995).
To reduce BVI noise levels, it is therefore essential to devise a control strategy
that targets the parameters affecting their intensity—namely, vortex strength and
vortex–blade separation distance. In this section two AFC strategies for control-
ling the aerodynamics of BVI are presented.

1. Blade-Mounted Trailing Edge Flap


Among the active BVI noise and vibration reduction techniques, the use of a
blade-mounted TE flaps (Fig. 3) has shown the most promise in both numerical
simulations (Straub and Hassan, 1996) and wind tunnel tests (Dawson et al.,
1995). The use of a TE flap for the alleviation of BVI noise is based on the premise

Fig. 3 The use of outboard TE flaps for the control of BVI noise and vibration:
a) photo of a four-bladed model rotor using TE flaps in the NASA Langley 14 ft ¥ 22 ft
wind tunnel; b) close-up photograph of the outboard flaps on the full-scale MD-900
helicopter rotor blades.
408 A. A. HASSAN ET AL.

that, by deploying the flap, one can alter the local aerodynamics of the blade and
thereby the associated vortex wake in terms of its strength and relative proximity
to the blades.
Figure 4 illustrates comparisons between the predicted temporal gradients of
the differential pressures (Cpu – Cpl) at the 1% chord position for the baseline and
for the controlled five-bladed MD-900 rotors at four radial positions, r/Rtip = 0.60,
0.70, 0.80, 0.90. The results are obtained using a coupled lifting-line free-wake
rotor integral code, CAMRAD.JA (Johnson, 1988), and the full potential FPRBVI
(Hassan, 1991) flow solver. Vortex-induced velocities due to the embedded free
wake were computed using the Biot-Savart law at grid point locations lying on the
surface of the blade. With a user-specified azimuthal flap deflection schedule, the
induced velocities due to the motion of the flap were also computed on the surface
of the flap.
The results of Fig. 4 are shown for an advance ratio of 0.22, a tip Mach number
of 0.625 and a tip path plane angle of –2 deg. Shown in Fig. 5 is the azimuth-de-
pendent (temporal) flap deployment schedule enforced between the 0.55 and 0.95
blade radial positions. The strong advancing BVI, associated with the higher local
Mach numbers, are seen to occur between the 25 deg and 100 deg blade azimuth
positions. By comparison, the relatively weaker retreating BVI associated with
lower Mach numbers are seen to occur between the 250 deg and 300 deg blade
azimuth positions. Figure 4 indicates that while the majority of the advancing BVI
LIVE GRAPH
Click here to view

Fig. 4 Predicted temporal pressure gradients near the LE of the baseline and the
controlled five-bladed MD-900 rotor. Mu = 0.22; Mtip = 0.625; TPP = -2 deg.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 409

Fig. 5 Simulated flap deployment schedule used in the numerical simulations.

are reduced in strength, only a few are reinforced. The intensity of the retreating
side BVI (represented by the peak-to-peak amplitude of the pressure gradients)
are essentially unaltered due to the zero flap deflection over the 180–360 deg
azimuth range. This figure suggests that more effective control of rotor BVI noise
can be achieved through the deployment of the flap over the entire blade azimuth
range of 0–360 deg (Charles et al., 1996).
For a baseline and a flapped four-bladed model rotor (Dawson et al., 1995),
Fig. 6 illustrates the measured noise levels during conditions representative of
LIVE GRAPH
Click here to view

Fig. 6 Effects of the TE flap on the noise footprint of a four-bladed model rotor
(Dawson et al., 1995).
410 A. A. HASSAN ET AL.

low-speed descent flight. The results are shown for an advance ratio of 0.1488, a
shaft angle of 5 deg aft and a normalized thrust coefficient (CT/s) of 0.0765. The
flap deployment schedule is identical to that shown in Fig. 5 with a –20 deg shift
in azimuth and a –12 deg peak deflection. The benefit of the TE flap on the advanc-
ing BVI noise levels is evident. The ineffectiveness of the flap on the retreating
BVI noise levels is a reflection of the flap deployment schedule of Fig. 5 where
only advancing BVI are being targeted.
Despite the demonstrated BVI noise reduction benefits in the experiments of
Dawson et al. (1995), an overall increase in the required rotor power was observed.
As with the deployment of all types of control surfaces, this increase in power is
expected and is a result of the increase in blade drag due to the deployment of the
TE flap. An AFC strategy that minimizes the drag penalty associated with the
deployment of the TE flap would therefore be desirable.

2. Reducing the Drag of a Plain Flap


Recently, Nishri and Wygnanski (1998) demonstrated that a low-momentum
oscillatory “zero-net-mass” jet of air can be used to postpone/prevent the separa-
tion of the boundary-layer flow over a simulated TE flap. Although the results
were limited to low Reynolds numbers, they demonstrated the potential benefits
of this novel fluidic AFC concept. For a practical rotor design, an “on-blade” AFC
method that emulates surface blowing without the actual transfer of mass would
be desirable since it eliminates the need for an air management system required to
provide air to the rotating blades.
The effectiveness of this fluidic AFC technique at Mach numbers and Reynolds
numbers typical of those for a rotor blade was numerically investigated by Hassan
(2003) for an idealized 2-D model problem employing the NACA-0012 airfoil
having a 1 ft chord and a 20% chord plain TE flap deflected 40 deg. The numerical
simulations, based on the solutions to the unsteady, 2-D, compressible Navier–Stokes
equations (Rumsey et al., 1996), employed a surface transpiration boundary condi-
tion to represent the temporal variation of the jet velocity. In this problem, the jet is
introduced at a point located on the upper surface of the flap at a distance of 0.72 c.
Jet width is equal to 0.0045 c, jet injection angle is 25 deg relative to the local
surface tangent and jet oscillation frequency is equal to 156 Hz (i.e., F + = 0.28).
The predicted streamwise component of velocity (u-component), as a function
of peak jet Mach number (Mjet = 0.0–0.30 or Cm = 0.0–8.1%) for the baseline and
for the controlled flapped NACA-0012 airfoils at a free-stream Mach number of
0.10, an angle of attack of 0 deg, and a Reynolds number of 1 million are shown
in Fig. 7a. Note the gradual reduction in the size of the recirculation flow regions
(see arrows on figure) with the increase in the jet peak Mach number. This reduc-
tion in size is a direct result of the favorable pressure gradients created over the
flap due to the control.
For a free-stream Mach number of 0.30 (more representative of the flow
environment over a flap on a blade) the predicted instantaneous streamwise
components of velocity (u-component), as a function of peak jet Mach number
(Mjet = 0.0–0.30 or, Cm = 0.0–0.9%, are shown in Fig. 7b. Again, for the con-
trolled airfoil, the jet oscillation frequency is 156 Hz (F + = 0.10). Unlike the
flowfields shown in Fig. 7a, only small reductions in the streamwise extent of
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 411

Fig. 7 Effects of introducing periodic packets of vorticity near the LE of the flap on
the flowfield. Shown are instantaneous constant u-velocity lines over the NACA-0012
airfoil as a function of free-stream Mach number and the intensity of the control
(Hassan, 2003).

the recirculation flow region on the upper surface of the flap are achieved with
the increase in the jet peak Mach number (note that Cm has been reduced by a
factor of 9 and the F + by a factor of 3).
Figure 8 illustrates the predicted percentage changes in airfoil lift and drag as a
function of free-stream angle of attack, free-stream Mach number and the intensity
of the control, Mjet. Clearly, an increase in actuator authority is desirable to maxi-
mize the aerodynamic benefit—be it a reduction in drag and/or an increase in lift.
Today, in the absence of high authority actuators capable of producing oscilla-
tory jets with significantly higher peak jet Mach numbers, it is expected that small
reductions in rotor power due to the deflection of the flap can be achieved with
AFC. While the observed benefits in airfoil lift would translate into enhanced rotor
thrust and therefore favor the use of this AFC strategy, the limited benefit in reduced
rotor power would, however, not warrant the added rotor design complexity
requiring the integration of an actuator inside the flap or blade. These two reasons
provide the impetus for the development of new high-authority actuator designs.

3. Use of Steady Blowing and/or Suction


In Sec. 1, the benefits of using a blade-mounted TE flap for the alleviation of
rotor BVI was discussed. The geometric parameters of the blade (or the airfoils
which constitute the blade) are, however, equally important in affecting the overall
intensity of the interactions. Most important are: a) the airfoil’s LE radius, b)
maximum thickness, c) maximum camber, d) the positions of maximum thickness
and maximum camber, and e) the thickness and camber distributions. These
parameters need to be optimized in order to minimize the airfoil’s response to the
interaction without resulting in poor aerodynamic performance at other flight
412 LIVE GRAPH A. A. HASSAN ET AL. LIVE GRAPH
Click here to view Click here to view

Fig. 8 Predicted percentage changes in the lift and drag forces of the flapped NACA-
0012 airfoil as a function of free-stream angle of attack, free stream Mach number
and the intensity of the control (Hassan, 2003).
LIVE GRAPH
LIVE GRAPH Click here to view
Click here to view
conditions. For example, improved transonic aerodynamic characteristics can be
achieved using an airfoil section with a maximum camber that is located far aft of
the LE and a small LE radius. However, a small LE radius will also guarantee poor
maximum lift characteristics at low speeds and high angles of attack which are
typical of the flow conditions on the retreating side of the rotor disk. Therefore, an
optimum airfoil/blade configuration would be one which involves a “compliant”
surface to meet the conflicting BVI requirements, the transonic high-speed flow
requirements, and the low-speed high-lift requirements.
For the control of BVI, hence BVI noise, an alternate strategy is one that utilizes
steady normal suction and/or blowing (Fig. 9). The primary objective is to alter
the temporal pressure gradients near the blade’s LE using a steady blowing jet, a
steady suction jet or a combination of steady blowing/suction jets on opposing
surfaces of the blade. Note that with the application of normal surface blowing/
suction, not only can one alter the effective camber of the airfoil, but one can also
directly influence its effective thickness as well as the chordwise position of maxi-
mum thickness. In this context, the word “effective” surface refers to the edge of
the boundary-layer displacement surface or to the bounding streamline beyond
which viscous flow effects are no longer dominant. For example, with the applica-
tion of blowing on the upper surface and suction on the lower surface of an other-
wise symmetrical airfoil one can emulate the same aerodynamic effects which
result from the use of a cambered airfoil section. A rotor blade (airfoil) that utilizes
normal surface blowing and/or suction would undoubtedly fulfill the above
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 413

Fig. 9 Sketch illustrating the use of upper/lower steady surface blowing/suction on


an MD-900 rotor blade (First appeared in Journal of the American Helicopter Society,
Hassan et al., 1996. Copyright AHS, reprinted with permission).

conflicting flow requirements providing, of course, that both thickness and camber
are temporally (i.e., as a function of blade azimuth) adjusted simultaneously to
meet the particular flight condition.
For the controlled MD-900 rotor blade, Fig. 10 illustrates three blowing/suction
“schedules” that represent the variation of the nondimensional blowing/suction
intensity, Vn, as a function of blade azimuth. Figure 9 indicates that suction/blow-
ing is applied between the LE of the blade and the 16% chord station for all blade
stations between 60% and the tip. The jet intensity Vn, also viewed as a jet Mach
number, is expressed as a fraction of the blade rotational tip Mach number. For
example, for Vn = 0.60 and for a blade tip Mach number of 0.6225, the peak jet
Mach number Vn is equal to 0.37. Figure 10 indicates that the application of
blowing and/or suction can be either azimuth-dependent (requiring a hub-based
regulator to control the flow of air around the azimuth) or, flight mode-dependent,
i.e., “on-off ” requiring only its activation during low-speed descent flight. In the
azimuth-dependent mode of operation, blowing/suction can be initiated to reduce
the intensity of the advancing BVI or the advancing blade shockwaves in high-
speed forward flight. In the flight mode-dependent operation, blowing can also be
maintained over the entire rotor disk to reduce the intensity of the less dominant
retreating BVI. This latter mode of operation, of course, requires a less complex
control system for administering the air through the rotating blades.
With blowing being applied over the entire rotor disk, it is expected that the
power requirements for this mode of operation (i.e., on-off) will be higher than
those for the azimuth-dependent mode of operation where blowing is carried out
only over the advancing side of the rotor disk. To quantify the power requirements,
one must revert to the use of a Navier–Stokes formulation in order to accurately
414 LIVE GRAPH A. A. HASSAN ET AL.
Click here to view

Fig. 10 Example 57 blowing/suction “schedules” representing the variation of the


non-dimensional blowing/suction intensity Vn as a function of blade azimuth.

predict the increase in rotor drag due to the normal momentum associated with the
air jet(s). In low-speed descent, however, power should not be of concern due to
the availability of excess engine power.
The effects of upper surface blowing during the advancing portion of the cycle
on the strength of the BVI at the 90% radial station is demonstrated in Fig. 11, by
Hassan et al. (1996). Plots are made of the predicted temporal gradients of the
differential pressures near the blade LE, x/c = 0.01, as a function of blade azimuth
and blowing intensity Vn. The results are obtained using the coupled integral free
wake/full potential formulation of Sec. 1 and are shown for a descent at a speed of
82 kt, an advance ratio of 0.20 and a tip Mach number of 0.6225. As the blowing
intensity is increased from a value of 0.30 to a value of 0.60, the peak-to-peak
amplitudes of the various advancing blade interactions are reduced. Since the
intensity of BVI noise is directly related to the temporal gradients of the LE pres-
sures (see, for example, Caraddona et al., 1988), we infer that the BVI noise levels
associated with the blowing intensity of 0.60 are lower than those associated with
a blowing intensity of 0.30, and lower than those for the baseline rotor. On the
average, the magnitudes of the peak-to-peak amplitudes with normal blowing and
an intensity of 0.60 are about one-third to one-half the values for the baseline rotor.
Similarly, Fig. 12 indicates that upper surface blowing can significantly reduce the
strength of the retreating side BVI—hence, the associated noise levels.
The results of Figs. 11 and 12 suggest that a blowing intensity Vn equal to 0.70
or 0.75 may further reduce the intensity of the BVI. However, caution should be
exercised as the increase in blowing intensity directly translates into a higher jet
Mach number. A blowing intensity of 0.70 translates into a jet Mach number
approximately equal to 0.44. With the correct combination of angle of attack, jet
Mach number, and freestream Mach number, the local flow on the blade may
LIVE GRAPH
Click here to view
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 415

Fig. 11 Predicted temporal pressure gradients for the baseline and the controlled
MD-900 rotor blades—advancing blade. Forward flight speed = 82 kt; Mu = 0.22;
Mtip = 0.6225 (Hassan et al., 1996) (Copyright permission granted by AHS).

become supersonic rather than the otherwise subsonic flow for the baseline blade.
In supersonic flow, one has to contend with the presence of shock waves, wave
drag and, of course, a new source of noise—high-speed impulsive (HSI) noise.
The upper bound on the blowing intensity must therefore be limited to subsonic
flow regime.
The effects of combined blowing on the upper surface and suction on the lower
surface (referred to here as BUSL) and vice versa, suction on the upper surface
and blowing on the lower surface (SUBL) on the intensity of the advancing BVI
LIVE GRAPH
Click here to view

Fig. 12 Predicted temporal pressure gradients for the baseline and the controlled
MD-900 rotor blades— —retreating blade. Forward flight speed = 82 kt; Mu = 0.22;
Mtip = 0.6225 (Hassan et al., 1996).
416 LIVE GRAPH A. A. HASSAN ET AL.
Click here to view

Fig. 13 Predicted temporal pressure gradients for the baseline and the controlled
MD-900 rotor blades—advancing blade. Forward flight speed = 82 kt; Mu = 0.22;
Mtip = 0.6225.

for the MD-900 rotor are shown in Fig. 13 for two-blade radial stations. For both
control strategies, identical intensities for blowing (Vn = 0.60) and suction
(Vn = -0.60) were simulated (Hassan et al., 1996) over the advancing side of the
rotor disk. In this example, no blowing and/or suction were simulated over the
retreating side. The results suggest that, overall, from a BVI alleviation perspec-
tive, the use of blowing on the upper surface combined with the use of suction on
the lower surface (i.e., BUSL) is slightly more advantageous than the use of suc-
tion and blowing on the upper and lower surfaces of the blade (i.e., SUBL) respec-
tively. As seen, the peak-to-peak values of the predicted temporal gradients are on
average 30–70% less than those observed for the baseline rotor, indicating a
milder response to the vortex encounters.

4. Emulating a Trailing Edge Flap Using a Transpiration Patch


Having shown the effectiveness of using a blade-mounted TE flap and normal
suction/blowing for reducing the strength of the advancing BVI, an obvious ques-
tion emerges–can the use of normal blowing/suction emulate the global aerody-
namic effects that result from the use of a TE flap? Specifically, is the impact of
blowing/suction on the rotor thrust and the trajectory of the tip vortex wake identi-
cal to that which results from the use of the TE flap? This question is answered
using results from a finite-element Euler flow solver (Dindar et al., 1999) for the
modified two-bladed, rectangular, untwisted Caradonna–Tung rotor (Caradonna
and Tung, 1981).
Figure 14a is a sketch illustrating the modeled blade which utilizes the NACA-
0012 airfoil. Here, steady normal blowing is emulated through the use of a trans-
piration boundary condition enforced over a user-specified patch on the lower
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 417

Fig. 14 The use of a transpiration patch on the lower surface of the Caradonna-
Tung blade to emulate the aerodynamic effects of a TE flap.

surface of the blade. As seen, the transpiration patch is bound by the 82% and 92%
radial positions and the 60% and 83% chord positions. Figure 14b depicts the
computed velocity vectors on the lower surface of the blade for a blade tip Mach
number of 0.439 and a collective pitch angle of 8 deg. In the simulation based on
the solutions to the Euler’s equations, the normalized transpiration velocity,
expressed as a fraction of the tip Mach number, is equal to 0.088.
Table 1 indicates that an increase in rotor thrust and the inviscid rotor power are
achieved by blowing. Alternately, note that if the surface patch is positioned on the
upper surface of the blade, then a decrease in rotor thrust and an increase in power
would be expected. This trend is identical to that observed for a helicopter blade
equipped with a plain TE flap. Figure 15a indicates that two secondary vortex
structures are created at the radial extremities of the surface transpiration patch.
These vortex lines mimic those created at both ends of a deployed finite span flap.
A downward shift in the position of the tip vortex wake is also observed when the
transpiration boundary condition is enforced on the lower surface of the blade
(Fig. 15b). This shift is equivalent to an increase in the average vortex–blade sepa-
ration distance—a factor deemed important for alleviating the impulsive aero-
dynamic response of the blade during BVI.

Table 1 Effects of the lower surface transpiration patch on the predicted thrust
and inviscid power for the two-bladed Caradonna–Tung rotor

Blade Mt Coli (deg.) CT Cq

Baseline 0.439 8 0.004601 0.00025


With control 0.439 8 0.004807 0.00042
Mtip = 0.439; collective pitch angle = 8 deg (Dindar and Hassan, 1999).
418 A. A. HASSAN ET AL.

Fig. 15 Use of a transpiration patch mimics the aerodynamic effects that result from
the use of a blade-mounted TE flap for the alleviation of BVI. Mtip = 0.439; collective
pitch angle = 8 deg (Dindar et al., 1999).

B. High-Speed Impulsive Noise—Control of Shock Waves


Under certain flight conditions, commercial and military helicopters/tiltrotors
generate an impulsive noise signature which is commonly referred to as HSI noise.
Specifically, in high-speed forward flight (or edgewise flight for tiltrotors), HSI
noise is generated from the advancing blades where local region(s) or pockets of
supersonic flow occur. At this flight condition, the supersonic flow region is usually
terminated with a strong shock wave, having a strength and attendant wave drag
proportional to the static pressure rise across the shock. On the rotating blade, the
strength of the shocks as well as the chordwise extent of the supersonic flow pocket
vary with azimuth due to variations in the local free-stream Mach number and
inflow angle. In hover, HSI noise can also occur due to a high rotational tip Mach
number, the combination of a moderate tip Mach number and a relatively thick
blade tip airfoil, or a moderate tip Mach number and a relatively thin tip airfoil that
is highly twisted. Regardless of the flight mode, however, once strong shock waves
have formed, they generate HSI noise that can severely impact the operation of
military and, to a lesser extent, commercial rotorcraft. An AFC strategy that mini-
mizes the strength of a shock (hence resulting in lower wave drag, lower potential
for shock-induced boundary-layer separation) is therefore desirable.
Extensive experimental and numerical investigations conducted recently by
Babinski and Ogawa (2006) and Smith et al. (2003, 2004) demonstrated that
reducing the strength of a shockwave, hence alleviating HSI noise and SBLIs, can
be effectively accomplished with the careful placement of 2-D/3-D surface bumps
or streamwise slots ahead of the shock. Results from their investigations have
shown that the drag reduction benefit from the different passive SBLI control
strategies are a direct result of shockwave bifurcation where a two-shock system
(a weak oblique shock and a weak normal shock) replaced the otherwise single
strong normal shock.
In this section we present results from a numerical investigation by Hassan
et al. (2007) using a Navier–Stokes flow solver (Rumsey et al., 1996) to demon-
strate the benefits of using a “transverse” oscillatory zero-net-mass flux jet to
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 419

reduce the strength of a normal shock on the NACA-64A010 airfoil. Grid cluster-
ing on the upper surface of the airfoil was performed to accurately capture the
details of the interaction between the transverse jet, located at x/c = 0.53, and the
external flow. Non-dimensional jet width, h/c, is equal to 0.0044. Figure 16 is a
sketch illustrating a number of states (A, a, B, b, C, c, D, . . .) during one jet oscil-
lation cycle. Noteworthy in this sketch are the states defined by the letters C and
G, which correspond, respectively, to the instants of peak suction and peak blow-
ing where the jet peak Mach numbers are equal to 0.30.
For a free-stream Mach number of 0.85, an angle of attack of 1 deg, a Reynolds
number of 6.07 million and a jet oscillation frequency of 123 Hz, Fig. 17 illustrates
comparisons between the predicted upper surface pressures during the instant of
peak suction C and peak blowing G. The predicted pressures for the baseline airfoil
are superimposed on the plot. Here, the lower surface pressures are not shown due
to their almost identical shapes for the blowing and suction extremes of the jet
oscillation cycle. Figure 17 suggests that the predicted 34.5% reduction in airfoil
lift is primarily attributed to the upstream movement of the upper surface shock due
to the control and, to a lesser extent, the very localized instantaneous changes in the
pressures in the vicinity of the jet during peak blowing.
Close-up views of the predicted upper surface temporal pressure distributions in
the vicinity of the jet and in the vicinity of the shockwave at the instants corre-
sponding to the letters A to I in Fig. 16 are shown in Figs. 18 and 19, respectively.
During the blowing portion of the jet oscillation cycle, Fig. 18 suggests that
airfoil lift is reduced. Here, the maximum reduction is seen to occur at point H or
I and not point G where peak blowing occurs. This lag reflects the time required
for the boundary-layer flow to adjust to the introduced blowing jet. Conversely,
during the suction portion of the jet oscillation cycle, the figure suggests that air-
foil lift is increased. Unlike the time lag associated with peak blowing, the effects
of peak suction are immediately felt at point C.
During the instant of peak blowing, Fig. 19 indicates that the upper surface
shock moves upstream to approximately the 64% chord position (vs the 67% posi-
tion for the baseline airfoil). On the other hand, during the instant of peak suction,
it is seen that the upper surface shock continues to move further upstream to the
LIVE GRAPH
Click here to view

Fig. 16 Simulated temporal variation of the intensity of the transverse jet. Note in
this example that the peak Mach number (or amplitude) of the jet is equal to 0.30.
420 A. A. HASSAN ET AL. LIVE GRAPH
Click here to view

Fig. 17 Predicted instantaneous pressure distributions for the NACA-64A010 airfoil


during peak blowing and peak suction. M = 0.85; a = 1 deg; Re = 6.07 million; jet
x/c = 0.53; jet oscillation frequency F = 123 Hz; slot width h/c = 0.0044; Mjet = 0.30;
jet injection angle = 90 deg (Hassan et al., 2007).

63.5% chord position. Therefore, in one jet oscillation cycle, the predicted
decrease in the “mean” lift of the airfoil is primarily due to the upstream unsteady
motion of the upper surface shock. Figure 19 indicates that the forward and aft
motions of the shock take place over approximately 0.5% of the chord during one
jet oscillation cycle. In general, the decrease in the “mean” drag of the airfoil

LIVE GRAPH
Click here to view
Fig. 18 Close-up view of the predicted instantaneous pressure distributions in the
vicinity of the unsteady transverse jet. M = 0.85; a = 1 deg; Re = 6.07 million; jet
x/c = 0.53; jet amplitude Mjet = 0.30; jet oscillation frequency F = 123 Hz; slot width
h/c = 0.0044; jet injection angle = 90 deg (Hassan et al., 2007).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 421

LIVE GRAPH
Click here to view
Fig. 19 Close-up view of the predicted instantaneous pressure distributions in the
vicinity of the unsteady transverse jet. M = 0.85; a = 1 deg; Re = 6.07 million; jet
x/c = 0.53; jet amplitude Mjet = 0.30; jet oscillation frequency F = 123 Hz; slot width
h/c = 0.0044; jet injection angle = 90 deg (Hassan et al., 2007).

during one jet oscillation cycle is seen to be a result of the more gradual pressure
rise (or alternatively, lower pressure gradients) across the shock, thereby leading
to a lower value of the dominant wave component of drag.
The effects of increasing the amplitude of the transverse jet at x/c = 0.53 on the
predicted upper surface pressures during the instant of peak blowing, point G is
shown in Fig. 20. With the increase in Mjet, three changes to the baseline airfoil
pressure distribution are seen. Namely, a) the upstream motion of the shock from
approximately x/c ~ 0.67 to x/c ~ 0.63, b) the gradual decrease in the pressure
gradient, d(Cp)/d(x/c), across the shock, and c) the increase in the upstream influ-
ence of the injected jet (emulating that of a weak compression wave). The results
of Fig. 20 suggest that airfoil mean lift is inversely proportional to the amplitude
of the jet (note the decrease in the area under the Cp–x/c curve). This figure also
suggests that the airfoil’s wave component of drag is inversely proportional to the
amplitude of the jet (note the decrease in the pressure gradients across the
shock).
To gain insight into the reason(s) for the reduction of airfoil mean drag, we
examine the details of the flow in the immediate vicinity of the jet. Figure 21
illustrates snapshots of the predicted constant Mach number contours in the vicin-
ity of the jet at selected instants (A, a, B, b, C, c, D, d, E) during the suction portion
of the jet oscillation cycle. For completeness, the predicted Mach number contours
for the baseline airfoil are also shown.
At point A, very close to the jet, Fig. 21 indicates the presence of a small
“bump-like” region associated with low Mach numbers (shown in green color).
This region is a remnant from the effects of the weak blowing jet at point I. In
general, for the time instants that follow point A, it is seen that a very thin bound-
ary layer results during the suction portion of the jet oscillation cycle. Moreover,
422 A. A. HASSAN ET AL. LIVE GRAPH
Click here to view

Fig. 20 Close-up view of the predicted instantaneous pressure distributions in the


vicinity of the upper surface shock as a function of peak Jet Mach number. M = 0.85;
a = 1 deg; Re = 6.07 million; jet x/c = 0.53; jet oscillation frequency F = 123 Hz; slot
width h/c = 0.0044; jet injection angle = 90 deg (Hassan et al., 2007).

it is seen that the largest influence of the jet on the external flow occurs at point C
where maximum suction takes place.
To explain the presence of compression waves in the vicinity of the temporal
jet, we refer to Fig. 22. At the jet injection position (x/c = 0.53) close examina-
tion of the figure reveals the evolution of a bump-like flow region whose dimen-
sions, in terms of chord extent and height, vary in direct proportion to the intensity
of the jet. Having reached a critical height, it is seen that compression waves are
formed on the upstream side of the bump—akin to those forming on the upstream
face of a bump (or a compression ramp) in supersonic flow. These waves, for jet
intensities approaching the peak intensity of 0.30, coalesce to form a weak
oblique shock that is visible at points F, f, G, g and H. These oblique shocks, in
turn, interact with the now weak normal shockwave, Fig. 17, located downstream
from the “fluidic bump.” Overall, due to the control, the sum of the wave drag
associated with the formation of the oblique shock and the weak normal shock is
approximately 4% less than that associated with the strong normal shock for the
baseline airfoil.
Figure 23 illustrates four schlieren-captured photos of the flowfields in the
experiments by Babinsky and Ogawa (2006) and Smith et al. (2003, 2004). In
their experiments, at a free-stream Mach number of 1.29, they demonstrated the
use of three different passive flow control strategies for reducing the strength of a
strong normal shock. Namely, a 3-D surface bump, a streamwise groove and a
slot/plenum. Shown also in the figure is the schlieren-captured photo of the formed
normal shockwave for the baseline wind tunnel setup. Despite the different control
strategies, striking similarities are seen in the captured flowfields. Specifically, the
formation of oblique shocks at the locations where the disturbances are introduced
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 423

Fig. 21 Close-up views of the predicted instantaneous constant Mach number con-
tours for the NACA-64A010 airfoil during the suction portion of the jet oscillation
cycle. M = 0.85; a = 1 deg; Re = 6.07 million; jet amplitude: Mjet = 0.30; jet oscillation
frequency: F = 123 Hz; jet injection position; x/c = 0.53; slot width: h/c = 0.0044; jet
injection angle = 90 deg (Hassan et al., 2007).

and, from a more global perspective, the bifurcation of the original single normal
shock into two weaker shocks.
Despite the differences between the freestream Mach number examined in the
investigations of Babinsky, Ogawa and Smith and that simulated in the investiga-
tion by Hassan et al. (2007), it is seen that the global features of the flow are
identical with the use of the transverse oscillatory jet. Here, of course, it is the
creation of a dynamic fluidic bump that is responsible for the bifurcation of the
otherwise strong shock and the reduction in the drag of the airfoil.
424 A. A. HASSAN ET AL.

Fig. 22 Close-up views of the predicted instantaneous constant Mach number con-
tours for the NACA-64A010 airfoil during the blowing portion of the jet oscillation
cycle. M = 0.85; a = 1 deg; Re = 6.07 million; jet amplitude: Mjet = 0.30; jet oscillation
frequency: F = 123 Hz; jet injection position: x/c = 0.53; slot width: h/c = 0.0044 jet
injection angle = 90 deg (Hassan et al., 2007).

C. Oscillatory Jets for Stall/Post-Stall Lift Enhancement


Increasing the maximum thrust capability, hence payload, remains of prime
importance to future military and commercial rotorcraft. Recent results (Greenblatt
and Wygnanski, 1998; Greenblatt et al., 1998; Nagib et al., 2001; Seifert et al.,
1996) from static and dynamic 2-D wind tunnel tests for the VR-7 and NACA-
0015 airfoils (used on the MH-47 Chinook, MD-500 series helicopter rotors),
supported by results from numerical simulations (Hassan, 2004), have shown that
an oscillatory zero-net-mass jet can be used to increase post-stall lift while
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 425

Fig. 23 Schlieren-captured photos illustrating the effects of three passive shockwave


boundary-layer interaction control strategies. (Figures courtesy of Prof. H. Babinsky,
Cambridge University.)

simultaneously reducing drag and pitching moments. In contrast to the use of


steady blowing or steady suction, this AFC strategy is based on the fact that by
introducing a low-momentum oscillatory jet at a critical point in the separated
boundary layer, one can partially, or fully, reattach the flow—thereby resulting in
the enhancement of lift and/or the reduction of drag (Bar-Sever, 1989).
While the transition of this AFC technology to advanced helicopter/tiltrotor
rotor systems demands mature, high-authority, airworthy actuators, clear under-
standing of practical actuator-blade integration issues is essential for the successful
implementation of this technology. For example, issues such as internal blade
packaging (Fig. 24), blade structural loads due to the added weight of the actuator,

Fig. 24 Conceptual integration of an electromagnetic oscillatory jet actuator inside


an airfoil section of a blade (US patents: 5,938,404 and 6,092,990).
426 A. A. HASSAN ET AL.

and the reliability/performance of the actuator under representative blade loading


are among the many factors that must be considered in the design of an advanced
rotor system that utilizes AFC. With packaging space being at a premium in a typi-
cal rotor blade, maximizing the attained aerodynamic benefits from an embedded
actuator becomes of prime importance. These benefits can result from either the
use of a more sophisticated actuator design or the use of an existing actuator with
an alternate AFC strategy that maximizes these benefits. This section addresses
the latter approach for maximizing the attained aerodynamic benefits.
In an earlier numerical investigation by Hassan (2004), it was shown that the
use of a pulsed suction jet to improve the post-stall aerodynamics of the VR-7
airfoil was more advantageous than the use of a zero-net mass oscillatory jet. It
was also demonstrated that the use of a pulsed blowing jet, when introduced at
the same free-stream conditions and location over the airfoil was, by comparison,
less effective and sometimes detrimental. Based on these results, an intriguing
question was raised. Namely, whether the simultaneous use of a pulsed suction
jet and a pulsed blowing jet can result in aerodynamic benefits that not only
exceed those obtained from the use of a pulsed suction jet but also from the use
of the oscillatory jet?
Figure 25 provides a new look at the general features of a two-port oscillatory jet
actuator. Also shown is a sketch of the traditional one-port zero-net-mass oscilla-
tory jet actuator discussed earlier in Chapter 6. Note that the oscillatory jet actuator
can also be viewed as one that has two orifice ports where each port is equipped
with a one-way valve to regulate the flow into and out of the actuator cavity. These
valves can be operated passively—for example, relying on changes in the internal
cavity pressure, or actively using an electronic solenoid. The two-port actuator of
Fig. 25 results in a pulsed suction jet and a pulsed blowing jet that are 180 deg out

Fig. 25 A “new look” at the features of a zero-net mass oscillatory jet actuator.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 427

of phase. However, as a whole, it remains an autonomous device that does not


require under-surface plumbing. With two orifice ports rather than one, the natural
question that follows is-where along the surfaces of the airfoil does one introduce
the two pulsed blowing and suction jets such that the resulting aerodynamic benefits
for an airfoil exceed those obtained from the use of either the pulsed suction jet or
the oscillatory jet?
Table 2 summarizes the jet injection positions on the upper surface of the VR-7
airfoil for the one-point and two-point AFC cases (Hassan, 2004, 2006). The table
entries under the headings ZNMJ (zero-net-mass jet), PS (pulsed suction) and, PB
(pulsed blowing) depict the dimensionless chord position(s) x/c for the jet injec-
tion point(s). The results presented are for a freestream Mach number of 0.30, an
angle of attack of 15 deg (airfoil static stall angle is 13 deg), a Reynolds number
of 2.14 million (based on an airfoil chord length of 1 ft) and, a jet width-to-chord
ratio, h/c, of 0.0035. For the controlled airfoils, the peak Mach number of the jet
is 0.30 (Cm = 0.70%) and the jet excitation frequency is 350 Hz (F + = 0.73 for
one-point control and for two-point control: F + = 0.73 for the upstream jet, 0.27
for the downstream jet).
Figure 26 illustrates comparisons between the predicted temporal lift coeffi-
cients when using the oscillatory jet, the pulsed suction jet, the pulsed blowing jet
and the two-point hybrid (case 2) pulsed suction/pulsed blowing jets. For the
hybrid AFC strategy, Table 2 indicates that in addition to the pulsed suction jet at
x/c = 0.30, a pulsed blowing jet is also introduced further downstream at x/c = 0.74.
As with the one-point control strategies, the two jets are assumed to be at angles
of 25 deg relative to the local surface tangents. Excitation frequencies for both jets
is kept constant at 350 Hz.
As seen, the two-point hybrid AFC strategy results in the maximum enhance-
ment in the post-stall sectional lift of the VR-7 airfoil—namely a 15% improve-
ment over the baseline airfoil value. By comparison, the use of pulsed suction,
considered the best among the three single-point AFC strategies, results in an
11% improvement over the baseline airfoil value. It is important to mention that
while the percent enhancements in airfoil sectional lift are not significantly dif-
ferent, the creation of a pulsed suction jet in the context of an “autonomous”
device is impossible.
The time-averaged mean sectional lift values shown in Fig. 26 indicate that
while an autonomous single port oscillatory jet results in a 7.3% enhancement in

Table 2 Matrix depicting the two airfoil chord positions (normalized by airfoil
chord length C) for introducing the oscillatory (ZNMJ), pulsed suction (PS) and
pulsed blowing (PB) jets

AFC Strategy Case no. ZNMJ PS PB

One-point a 0.30 na na
One-point b na 0.30 na
One-point c na na 0.30
Hybrid (two-point) 2 na 0.30 0.74
With the exception of cases a, c, note that pulsed suction is always applied at x/c = 0.30.
428 A. A. HASSAN ET AL. LIVE GRAPH
Click here to view

Fig. 26 Predicted temporal lift coefficients illustrating the benefit of the hybrid suc-
tion/blowing technique over those obtained using pulse suction, pulse blowing or an
oscillatory jet. Mjet = 0.30; F = 350 Hz; jet angle = 25 deg.

post-stall lift, a two-port autonomous device producing pulsed suction/blowing


jets can be used to achieve a 15% enhancement—in essence, doubling the percent
benefit achieved with the use of the single-port oscillatory jet. It is also seen that
the use of the pulsed blowing jet as the primary means to enhance VR-7 airfoil
post-stall lift is detrimental. Yet, when used in conjunction with the pulsed suction
jet in a two-point hybrid AFC strategy, then the maximum enhancement in VR-7
airfoil post-stall lift is achieved.
Comparisons between the predicted temporal sectional drag values are shown
for the one- and two-point AFC strategies in Fig. 27. It is seen that the hybrid
control strategy results in the largest percent reduction, 32%, in the drag value of
the baseline airfoil. By comparison, the use of the one-port oscillatory jet results
in a 12% reduction in post-stall drag. Again, it is clear from the perspective of an
autonomous AFC device, that the use of the two-point hybrid AFC strategy results
in drag reduction benefits that are more than twice those achieved using the single-
port oscillatory jet.
The predicted time histories of the pitching moment coefficients for the VR-7
airfoil for the one- and two-point hybrid AFC strategies are shown in Fig. 28. As
seen, the lowest mean pitching moment value (in absolute sense: 0.0195) is
achieved using the two-point case-2 hybrid AFC strategy. This value represents a
33% reduction in the post-stall pitching moment of the baseline VR-7 airfoil. By
comparison, for the one-point control strategy utilizing the one-port oscillatory
jet, case-a, it is seen that the mean pitching moment is equal to -0.0281 represent-
ing, approximately, a 3% reduction in the pitching moment of the baseline airfoil.
In this respect, the two-point hybrid AFC strategy provides pitching moment
reduction benefits that are more than 10 times those achieved using the AFC
strategy utilizing the one-port oscillatory jet. This benefit is attributed to the
changes in the local surface pressures near x/c = 0.74 where the pulsed blowing jet
is introduced.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 429

LIVE GRAPH
Click here to view
Fig. 27 Predicted temporal drag coefficients illustrating the benefit of the hybrid
suction/blowing technique over those obtained using pulse suction, pulse blowing or
an oscillatory jet. Mjet = 0.30; F = 350 Hz; jet angle = 25 deg.

A close-up view of the time-averaged predicted particle traces near the TE of the
VR-7 airfoil when only pulsed suction is used is shown in Fig. 29. For contrast, the
predicted particle traces for the baseline airfoil are also shown. For the two-point
hybrid AFC case, close-up views of the predicted instantaneous particle traces in
the vicinity of the airfoil’s TE are shown in Fig. 30. Also shown in the figure is a
sketch depicting the variation of the instantaneous jet velocity (Mach number) at

LIVE GRAPH
Click here to view
Fig. 28 Predicted temporal pitching moment coefficients illustrating the benefit of
the hybrid suction/blowing technique over those obtained using pulse suction, pulse
blowing, or an oscillatory jet. Mjet = 0.30; F = 350 Hz; jet angle = 25 deg.
430 A. A. HASSAN ET AL.

Fig. 29 Close-up views of the predicted instantaneous particle traces for the baseline
and the controlled VR-7 airfoils. M = 0.30; a = 15 deg; Re = 2.14 million; xj /c = 0.30;
Mjet = 0.30; F = 350 Hz; jet angle = 25 deg (Hassan, 2006).

various states a–h. A comparison between the particle traces from the one-point
control strategy using pulse suction and the two-point hybrid strategy reveals that
the introduction of the pulsed blowing jet has changed the features of the local flow
in the vicinity of the airfoil’s trailing edge. Namely, fragmenting and reducing the
size of the otherwise relatively larger (in chord extent, height) separated flow region
that existed near the TE when only pulsed suction was used.
During the active period of the pulsed blowing jet, and especially near peak
blowing (points e, f, g), it is seen that the fluid particles are drawn closer towards
the surface of the airfoil. This favorable effect is a result of imparting additional
momentum to the otherwise low-energy separated boundary-layer flow. As a
result, lower surface pressures that draw the fluid closer to the surface of the airfoil
are created in the vicinity of the jet. During the dwell period (i.e., points a, b, c),
the local flow gradually regains its original state where a recirculation flow region
existed over this extent of the airfoil chord.
In Fig. 30 it is observed that large portions of the boundary-layer flow are drawn
closer to the surface of the airfoil despite the fact that the instantaneous jet Mach
numbers at points g and h are gradually approaching zero. This behavior reflects

Fig. 30 Close-up views of the predicted instantaneous particle traces for the twopoint
hybrid AFC strategy for the VR-7 airfoil. M = 0.30; a = 15 deg; Re = 2.14 million;
Mjet = 0.30; F = 350 Hz; jet angle = 25 deg; suction: xj /c = 0.30; blowing: xj /c = 0.74.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 431

the time lag between the instant where the jet emanates from the surface slot
and the time required for the boundary-layer flow to adjust to this local distur-
bance. In the same figures, we note the gradual increase in the size (chord extent)
of the recirculating flow regions just upstream of the airfoil’s TE. This growth in
size is reflective of the increase in the adverse local pressure gradient that is expe-
rienced as the intensity of the jet approaches zero and remains at zero during the
dwell period (i.e., points a, b, c) of the cycle.
For multi-point AFC strategies, the results described above suggest that the
maximum attainable benefits in airfoil lift, drag, and pitching moment can be
arrived at using an optimization scheme that is coupled with the flow solver. The
results presented here were obtained using a laborious trial and error approach.

D. XV-15 Tiltrotor Hover Download Reduction


Tiltrotor aircraft such as the Bell XV-15 and the Bell/Boeing V-22 Osprey are
revolutionary twin-rotor vehicles that combine the low-speed benefits of helicop-
ters and the high-speed cruise benefits of a fixed-wing turbo-prop aircraft,
(Fig. 31). In hover, the wakes from the two rotor systems impinge on the upper
surface of the wing resulting in what is commonly referred to as a download force.
This force, Cn, also regarded as vertical drag since it is aligned with the direction
of the impinging flow, is known to significantly impact on the aerodynamic per-
formance of the vehicle and, ultimately, its useful payload. For the XV-15 aircraft,
this penalty in payload can be as high as 15% of the gross weight if the wing flaps
are not deflected, and as low as 10% of the gross weight with the outboard flaps
deflected downwards at 60 deg to reduce the projected wing area (Maisel et al.,
1986). This roughly represents the total useful load carried by the fully fueled
aircraft. For larger flap deflection angles, the boundary layer separates on the flap
resulting in an increase in the download.

1. Numerical Simulations
Consider the simplified 2-D model problem of a uniform free-stream impinging
on the upper surface of a modified NACA-64A223 airfoil (for geometry of the

Fig. 31 Photograph of a) the V-22 and b) the XV-15 tiltrotor aircraft in hover.
432 A. A. HASSAN ET AL.

cross-section of the flapped wing on the XV-15 aircraft (see Maisel et al., 1975).
This simplification is a necessary step to understand the resulting complex flow
and, more importantly, to identify the most effective control requirements for
successful reduction of the vertical drag.
The logic behind the use of AFC to reduce the download is easily explained by
referring to Fig. 32, which depicts a close-up view of the computed instantaneous
streaklines in the vicinity of the baseline airfoil. In the numerical simulations by
Roth (2003) for the flow past a circular cylinder, it was shown that an oscillatory
jet can be effectively used to reduce the drag. This reduction in drag and, in par-
ticular, the dominant pressure component of drag, was a direct result of reducing
the average width of the wake. Since the NACA-64A223 airfoil geometry can be
similarly viewed as a bluff body in crossflow, the primary objective from the use
of AFC was to reduce the average width of the wake below the airfoil from L to
L¢ by reattaching the separated boundary-layer flow over the flap.
Through a series of comprehensive wind tunnel tests and large eddy simula-
tions (LES), Kjellgren et al. (2002) demonstrated that a low-momentum oscilla-
tory zero-net-mass jet can be used to reduce the download (vertical drag) on the
2-D flapped NACA-64A223 airfoil placed normal to the free-stream. In their stud-
ies, emphasis was placed on exploring the effects of slot injection position, the
momentum of the jet, and the jet excitation frequency on the attained download
reduction benefits. When the near-tangent jet was introduced at a position slightly
downstream from the separation point, reattachment of the otherwise separated

Fig. 32 Streaklines depicting the instantaneous width of the wakes below the flapped
NACA-64A223 airfoil in the absence/presence of AFC. M = 0.138; a = -90 deg; Re = 1
million; d = 85 deg; flap chord = 0.25c (Hassan, 2004).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 433

Fig. 33 Computed instantaneous pressure contours depicting the vortex structures


in the flowfield. a) Baseline when the computed CD is a maximum; b) baseline pres-
sure field when the computed CD is at a minimum; and c) periodic excitation being
applied from a slot located at 10% of the flap’s chord.

boundary-layer flow over the flap was achieved and a 30% reduction in download
was demonstrated.
For an angle of attack of -85 deg and a flap deflection of 80 deg, Fig. 33 illus-
trates the computed instantaneous constant pressure contours for the baseline
airfoil when a) the computed CD is at a maximum, and b) when CD is at a mini-
mum. Also shown are predicted pressure contours c) when periodic excitation is
applied from a slot located at 10% of the flap chord.
When the oscillatory CD is close to its temporal maximum, the results reveal
the presence of a large vortex below the airfoil. This vortex is created by the rollup
of the mixing layer that separates from the LE of the airfoil. Here, the mixing
layer is susceptible to the Kelvin–Helmholtz instability that creates the smaller
array of vortices also visible below the LE and TE of this airfoil. When the instan-
taneous CD is at a minimum, the previously observed large circulation pattern is
seen to be swept further downstream, thereby minimizing the effects of the vortex-
induced velocity field on the lower surface of the airfoil. For the controlled airfoil,
in addition to the reattachment of the separated boundary-layer flow over the flap,
the figure suggests that a weak circulation pattern might have been created one
chord length below the airfoil. This pattern will have only a minor effect on the
pressure distribution over the airfoil and will reduce substantially the oscillations
in drag.
The time-averaged pressure distributions over the airfoil and flap for the base-
line and controlled cases are shown in Fig. 34. It is seen that the introduction of
the periodic jet over the flap has the following two effects: increasing the base
pressure on the lower surface of the airfoil, and substantially lowering the pres-
sure over the upper surface of the flap due to the reattachment of the flow over its
shoulder. These two factors are responsible for reducing the airfoil download.
Using an unsteady Reynolds-averaged Navier–Stokes flow solver (Rumsey
et al., 1996), an assessment was made by Hassan (2004) of the relative effective-
ness of two alternate AFC strategies for reducing the download on the modified
flapped NACA-64A223 airfoil. In addition to the use of an oscillatory jet, the
alternate control strategies encompassed the use of near-tangent pulsed blowing
434 A. A. HASSAN ET AL.

Fig. 34 Calculated and measured pressure coefficients over the baseline and the
controlled airfoil (a = -85 deg and d = 80 deg). a) Baseline flow; b) periodically excited
flow, from a slot located at 10% of flap chord (Kjellgren et al., 2002).

and pulsed suction jets. Similar to the numerical investigation by Kjellgren et al.
(2002), these new strategies targeted the reattachment of the separated flow on the
upper surface of the deflected flap and employed a transpiration boundary condi-
tion to emulate the temporal variations in the oscillating and pulsed jet velocities.
Figure 35 compares the predicted time histories of the vertical drag coefficients,
CD, for the flapped NACA-64A223 airfoil when an oscillatory jet, a pulsed blow-
ing jet, and a pulsed suction jet are introduced at x/c = 0.795 on the upper surface
shoulder of the flap. For the controlled airfoil, the jet peak Mach number is 0.138
(Cm = 2.268%) and the jet excitation frequency is 290 Hz (F + = 0.3933). The pre-
dicted mean drag for the baseline airfoil is shown as a dashed black line.
LIVE GRAPH
Click here to view

Fig. 35 Predicted time histories of the vertical drag coefficients, CD, for the flapped
NACA-64A223 airfoil as a function of the simulated AFC strategy. M = 0.138;
a = -90 deg; Re = 1 million; d = 85 deg; flap chord = 0.25c; F = 290 Hz (Hassan, 2004).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 435

Fig. 36 Predicted instantaneous particle traces for an oscillatory jet, a pulsed blow-
ing jet (peak blowing), and a pulsed suction jet (peak suction). M = 0.138; a = -90 deg;
Re = 1 million; d = 85 deg; flap chord = 0.25c; Mjet = 0.20; F = 290 Hz (Hassan, 2004).

The observed reductions in airfoil mean vertical drag can be explained by refer-
ring to Fig. 36 where snapshots of the predicted particle traces at three instants
representing a) peak suction for the oscillatory jet, b) peak blowing for the pulse
blowing jet, and c) peak suction for the pulse suction jet are shown.
Figure 36 indicates the presence of a “vortex train” with clockwise (CW) and
counterclockwise (CCW) vortices that convect towards the TE of the flap. The
CCW vortices result in acceleration of the local flow near the surface of the
flap—therefore resulting in a beneficial effect as they tend to locally reduce
the adverse pressure gradient. On the other hand, the CW vortices result in decel-
eration of the local flow near the surface of the flap—thereby resulting in further
increases in the adverse pressure gradient. As the vortices continue to propagate
along the surface of the flap, they increase in size due to their entrainment of the
external flow. Figure 36 suggests that, in order to promote further reattachment
of the flow over the flap, one has to increase the magnitude of the local suction
pressures or, equivalently, increase the strength of the CCW vortices. This is
accomplished by considering a higher intensity jet having a peak Mach number
of 0.25 (Cm = 3.54%), Fig. 37.
The computed percent reductions (Hassan, 2004) in the vertical drag of the
airfoil as a function of the peak jet Mach number and the simulated AFC strategy
is shown in Fig. 38. For a peak jet Mach number of 0.30, it is seen that the pulse
suction jet results in the maximum percent reduction of 32.1% in vertical drag
(cn). By comparison, the use of the oscillatory jet results in a 27.5% reduction in
436 A. A. HASSAN ET AL.

Fig. 37 Predicted instantaneous particle traces for an oscillatory jet, a pulsed blow-
ing jet (peak blowing), and a pulsed suction jet (peak suction). M = 0.138; a = -90 deg;
Re = 1 million; d = 85 deg; flap chord 0.25c; Mjet = 0.25; F = 290 Hz (Hassan, 2004).

vertical drag. The use of the pulse blowing jet is, however, shown to be ineffective
and, for certain jet intensities, can be detrimental.

2. Wind Tunnel Experiments and Flight Tests


AFC experiments were performed on a 16.7% scale powered model of the
XV-15 aircraft in the Aerodynamics Laboratory at the University of Arizona

LIVE GRAPH
Click here to view
Fig. 38 Summary of the predicted percent reduction in vertical drag for the flapped
NACA-64A223 airfoil as a function of the peak jet Mach number. M = 0.138; a =
-90 deg; Re = 1 million; d = 85 deg; flap chord = 0.25c; F = 290 Hz (Hassan, 2004).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 437

Fig. 39 Sketch and photo of the download rig for the 16.7% scale model of the XV-15
aircraft (Stalker, 2004).

(Stalker, 2004). The model was suspended from an A-frame structure in an inverted
arrangement, directing the rotor downwash up toward the ceiling, which was
located more than five rotor diameters away (Fig. 39). The A-frame structure
provided minimal interference to the resulting flow in the vicinity of the wing.
The drive system for the rotors was placed on the floor with gearing providing the
rotational motion up through two stacks to each of the rotor hubs. In this setup, the
rotors were not coupled to the nacelles to allow for a direct measurement of the
aerodynamic forces on the model. The gap between the rotors and the model’s
nacelles was less than 2 in. allowing for the placement of two force balance on top
of the rotor-hubs.
Five voice-coil actuators were internally mounted inside each wing. The actua-
tors provided oscillating zero-net-mass flux jets at a given location on the flap
surface. The periodic airflow was ducted from the wing to the flap and out through
a set of five slots having a width of 0.04 in. On the aircraft model, three slots were
located on the aileron and two along the inboard flap. A sinusoidal command
signal powered all the actuators that generated the zero mass flux forcing (ZMFF).
For the baseline and the controlled aircraft, Fig. 40 depicts the measured variation
of download as a function of flap-aileron angles at two rotor RPMs.
For a rotor RPM of 2200, a flap deflection angle of 60 deg and, a CT /s of 0.016,
the baseline DL / T was measured as 10.6%. With AFC turned on at Cm = 3.2%,
the minimum download was reduced by 13%. Reducing Cm to 1.25% only reduced
the download alleviation to 12%. For a rotor RPM of 1800, it is seen the DL/T has
been reduced by 16.6% for a Cm of 3.4%.
In June 2003, under the DARPA Micro Adaptive Flow Control (MAFC) pro-
gram*, the Bell XV-15 Tiltrotor aircraft (Fig. 41), was used for a series of flight tests
that demonstrated the effectiveness of AFC in reducing airframe download during
hover (Nagib et al., 2004). The aircraft (Fig. 31), was fitted with special-purpose

*Three universities (The University of Arizona, Tel Aviv University, and Illinois Institute of

Technology) and two major Aerospace companies (Bell Helicopters and the Boeing Company) were
involved in this program, administered by the U.S. Army Research Office under a contract from
DARPA. The authors wish to acknowledge the assistance of J. McMichael, R. Wlezien, and S. Walker
of DARPA, who navigated this program through its different phases to its successful conclusion.
438 LIVE GRAPH A. A. HASSAN ET AL.
Click here to view
XV-15 3D 16% Model with Internal Actuation
Study of Aileron and Flap Deflection with Full Span AFC
Slot @ 10%c
0.125

0.120 Baseline Untaped, 1800 RPM

0.115 3.4% Cmu, 1800 RPM

Baseline Untaped, 2200 RPM


0.110
3.2% Cmu, 2200 RPM

0.105
DL/T

1.25% Cmu, 2200 RPM

0.100
13.2%

0.095
16.6%

0.090

0.085

0.080
50 55 60 65 70 75 80 85 90
Aileron and Flap Angle

Fig. 40 Download reduction as a function of aileron and flap angles on the 16.7%
scale model of the XV-15 aircraft (Stalker, 2004).

flaps and ailerons fabricated by Bell Helicopters. Wing flaps were retrofitted with
zero-net-mass flux actuators that excited the flow over their upper surface by peri-
odically injecting/removing fluid through slots connected to cavities in their interior
(Fig. 39). Actuators used on the aircraft were manufactured and tested at the Illinois
Institute of Technology (Nagib et al., 2004). The flight tests were the culmination of
extensive 2-D and 3-D laboratory experiments that started by examining the effects
of AFC on an airfoil with its chord approximately normal to the impinging stream
(Kjellgren et al., 2000, 2002).
The performance goals of the program were: a) to demonstrate that ZMFF
could be used to reattach an otherwise fully-separated flow over the flap, and b) to
produce a useful reduction in the download by maintaining the flow attached
to the flap at higher flap deflection angles (Fig. 42). A reduction of 220 lb was

Fig. 41 The full-scale pseudo-flap and the actual design of the slot on the XV-15
flight test tiltrotor aircraft (Nagib et al., 2004).
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 439

Fig. 42 Flight test success criteria for the XV-15 tiltrotor aircraft (Nagib et al.,
2004).

targeted for goal a and 150 lb for goal b. All testing took place at Bell’s Arlington,
TX, flight test facility.
The aircraft was weighed with full fuel, and the weight of the pilots and their
baggage was also determined before each flight. A fuel flow totalizer monitored
the amount of fuel consumed and hence the weight at any time. Testing was con-
ducted early in the morning when the winds were less than 4 kt. The test procedure
was to hover out of ground effect at a carefully maintained altitude with a selected
flap setting with and without ZMFF. Data was taken first with the AFC system off,
then with the AFC system switched on, and finally with the AFC system switched
off again prior to changing the amplitude of the forcing. This procedure was
repeated for a range of flap settings, rotor rpm, and for a range of actuation param-
eters (Cm and F +).
The essential results are summarized in Fig. 43, which illustrates the variation
of rotor power coefficient Cp as a function of weight coefficient Cw in the presence
and absence of AFC. The results are shown for a flap deflection angle of 75∞
where the flow over the flap is fully separated for the baseline aircraft configura-
tion. The ability of AFC to reattach the flow over the flap is evident. For constant
power, the increase in payload (or Cw) due to AFC is greater than the target goal
of 220 lb—amounting to a 15% reduction in the aircraft download. When a com-
parison of the measured download values at the two minimum flap settings was
made, the results indicated that a reduction of 150 lb in download was also
achieved due to AFC, thereby meeting the second goal.

III. Prospects and Challenges of Modern Flow Control


Ten years ago the notion of using an oscillatory zero-net-mass flux jet for
actively controlling the flow in a separated boundary layer was perceived as being
440 LIVE GRAPH A. A. HASSAN ET AL.
Click here to view

Fig. 43 Flight test results illustrating the increase in the payload of the XV-15 air-
craft due to AFC (zero-net-mass flux jets) for a flap deflection angle of 75 deg (Nagib
et al., 2004).

a high-risk approach. Lacking were the fundamental understanding of the physi-


cal mechanisms responsible for the aerodynamic benefits and, more importantly,
actuators that had sufficient control authority to demonstrate the benefits at realis-
tic vehicle flight conditions. Today, armed with deep understanding of the physics,
the availability of more comprehensive prediction tools, an abundance of data
from a myriad of wind tunnel/flight tests, and advances in the development of
oscillatory jet actuators, it is no surprise that this particular AFC technology has
become the most widely recognized, not only in the rotorcraft industry but also in
the broad aerospace industry. The following statements reflect the current status of
AFC technology for rotorcraft:
1) Identification of the “correct” location(s) to apply the control on a airfoil/
blade/wing and/or a helicopter fuselage remain an essential factor for a successful
flow control strategy—irrespective of the capabilities of the actuator.
2) The benefits of an oscillatory jet for enhancing the stall/post-stall lift charac-
teristics of rotor blade airfoils have been demonstrated under various US Army
programs involving successful static/dynamic wind tunnel tests.
3) Reducing bluff body pressure drag of a helicopter fuselage and the wave
drag associated with strong shockwaves on the advancing rotor blade are two
potential beneficiaries of AFC technology.
4) Since space is at a premium, actuator/blade integration remains a challenge
for the implementation of AFC technology.
5) Oscillatory jet actuators with sufficient authority to control the flow phenom-
ena at realistic flight Mach numbers (<0.50) have matured beyond their early wind
tunnel predecessors. New designs are now necessary to meet the AFC require-
ments at transonic and supersonic Mach numbers.
FLOW CONTROL FOR ROTORCRAFT APPLICATIONS 441

6) A need remains for robust in-blade actuator designs that can withstand the
harsh rotor CF environment (up to 700 g).
7) Actuation requirements (e.g., excitation frequencies), while more demand-
ing for smaller scale model tests, will be significantly relaxed for large scale and
full scale configurations.
8) Actuator reliability (life cycle) has continued to improve. Additional design
improvements will be necessary to arrive at a “flightworthy” actuator.
9) Despite the demonstrated aerodynamic benefits, a need still exists to verify
the viability of AFC in a “rotating” flow environment.
10) Availability of accurate prediction tools and robust optimizers is essential in
the identification of the location(s) for actuation and the combination of actuation
parameters that yield the maximum aerodynamic benefits.
References

Aamo, O. M., Krstic, M., and Bewley, T. R. (2003), “Control of Mixing by Boundary
Feedback in 2D Channel Flow,” Automatica, Vol. 39, pp. 1597–1606.
Abdessemed, N., Sherwin, S., and Theofilis, V. (2004), “On Unstable 2D Basic States
in Low-Pressure Turbine Flows at Moderate Reynolds Numbers,” AIAA Paper
2004-2541.
Abdessemed, N., Sherwin, S., and Theofilis, V. (2006), “Linear Stability of the Flow Past a
Low-Pressure Turbine Blade,” AIAA Paper 2006-3530.
Abeyounis, W. K., and Bennett, B. D. Jr. (1997), “Static Internal Performance of an
Overexpanded, Fixed-Geometry, Nonaxisymmetric Nozzle with Fluidic Pitch-
Thrust-Vectoring Capability,” NASA TP-3645.
Abramson, J. (1977), “Two-Dimensional Subsonic Wind Tunnel Evaluation of Two Related
Cambered 15-Percent Circulation Control Airfoils,” DTNSRDC ASED-373.
Abramson, J., and Rogers, E. O. (1983), “High-Speed Characteristics of Circulation
Control Airfoils,” AIAA Paper 83-0265.
Ackeret, J., Betz, A., and Schrenk, O. (1926), “Experiments with an Airfoil from which the
Boundary Layer is Removed by Suction,” NACA-TM-374.
Agashe, J. S., Sheplak, M., Arnold, D. P., and Cattafesta, L. N. (2008), “MEMS-Based
Actuators for Flow-Control Applications,” IUTAM Symposium on Flow Control and
MEMS, edited by J. F. Morrison, D. M. Birch, and P. Lavoie, IUTAM Book Series,
Vol. 7, pp. 25–32.
Ahmed, S. R. (1983), “Influence of Base Slant on the Wake Structure and Drag of Road
Vehicles,” Journal of Fluids Engineering, Vol. 105, pp. 429–434.
Ahuja, K. K., and Brown, W. H. (1989), “Shear Flow Control by Mechanical Tabs,” AIAA
Paper 1989-0994.
Ahuja, K. K., and Burrin, R. H. (1984), “Control of Flow Separation by Sound,” AIAA
Paper 1984-2298.
Ahuja, K. K., Lepicovsky, J., and Burrin, R. H. (1982), “Noise and Flow Structure of a
Tone-Excited Jet,” AIAA Journal, Vol. 20, No. 12, pp. 1700–1706.
Ahuja, K. K., Whipkey, R. R., and Jones G. S. (1983), “Control of Turbulent Boundary
Layer Flow by Sound”, AIAA Paper 83-0726.
Airiau, C., Walther, S., and Bottaro, A. (2000), “Non-Parallel Receptivity and the Adjoint
PSE,” IUTAM Laminar-Turbulent Symposium V, edited by W. Saric and H. Fasel,
Sedona, AZ, pp. 57–62.

443
444 REFERENCES

Akervik, E., Hoepffner, J., Ehrenstein, U., and Henningson, D. S. (2007), “Optimal Growth,
Model Reduction and Control in a Separated Boundary-Layer Flow using Global
Eigenmodes,” Journal of Fluid Mechanics, Vol. 579, pp. 305–314.
Alam, M. R., Liu, W., and Haller, G. (2006), “Closed-Loop Separation Control: An Analytic
Approach,” Physics of Fluids, Vol. 18, 043601.
Alkislar, M. B., Krothapalli, A., and Butler, G. W. (2007), “The Effect of Streamwise
Vortices on the Aeroacoustics of a Mach 0.9 Jet,” Journal of Fluid Mechanics, Vol. 578,
pp. 139–169.
Allan, B. G., Juang, J. N., Raney, D. L., Seifert, A., Pack, L. G., and Brown, D. E. (2000),
“Closed-loop Separation Control Using Oscillatory Flow Excitation,” NASA/
CR-2000-210324.
Allan, B. G., Owens, L. R., and Lin, J. C. (2006), “Optimal Design of Passive Flow Control
for a Boundary-Layer-Ingesting Offset Inlet Using Design-of-Experiments,” AIAA
Paper 2006-1049.
Allen, M. G., Butler, S. A., Johnson, E. Y., and Russo, F. (1993), “An Imaging Neural
Network Combustion Control System for Utility Boiler Applications,” Combustion
and Flame, Vol. 94, pp. 205–214.
Alonso, J. (2007), “Overview of NRA Solicitation: N + 3 Pre-Proposal Conference,”
L’Enfant Plaza Hotel, Washington, DC., Nov. 29.
Alvi, F. S., Strykowski, P. J., Krothapalli, A., and Forliti, D. J. (2000), “Vectoring Thrust in
Multiaxes Using Confined Shear Layers,” Journal of Fluids Engineering, Vol. 122, No.
1, pp. 3–13.
Alvi, F. S., Shih, C., Elavarasan, R., Garg, G., and Krothapalli, A. (2003), “Control of
Supersonic Impinging Jet Flows Using Supersonic Microjets,” AIAA Journal, Vol. 41,
No. 7, pp. 1347–1355.
Amitay M., and Glezer, A. (2002), “Role of Actuation Frequency in Controlled Flow
Reattachment over a Stalled Airfoil”, AIAA Journal, Vol. 40, No. 2, pp. 209–216.
Amitay, M., Smith, B. L., and Glezer, A. (1998), “Aerodynamic Flow Control Using
Synthetic Jet Technology,” AIAA Paper 98-0208.
Amitay, M., Parekh, D., Pitt, D., Kibens, V., and Glezer, A. (2000), “Control of Internal
Flow Separation Using Synthetic Jet Actuators,” AIAA Paper 2000-0903.
Amitay, M., Smith, D. R., and Kibens, V. (2001), “Aerodynamic Flow Control over an
Unconventional Airfoil Using Synthetic Jet Actuators,” AIAA Journal, Vol. 39, No. 3,
pp. 361–370.
Amitay, M., Washburn, A. E., Anders, S. G., and Parekh, D. E. (2004), “Active Flow Control
on the Stingray Uninhabited Air Vehicle: Transient Behavior,” AIAA Journal, Vol. 42,
No. 11, pp. 2205–2215.
Anderson, B. D. O., and Moore, J. (1990), Optimal Control: Linear Quadratic Methods,
Prentice Hall, Englewood Cliffs, NJ.
Anderson, D. A., Tannehill, J. C., and Pletcher, R. H. (1984), Computational Fluid
Mechanics and Heat Transfer, McGraw-Hill, New York, p. 503.
Anderson, C. J., Giuliano, V. J., and Wing, David J. (1997), “Investigation of Hybrid
Fluidic/Mechanical Thrust Vectoring for Fixed-Exit Exhaust Nozzles,” AIAA Paper
97-3148.
Anderson, B. H., and Gibb, J. (1998), “Vortex Generator Installation Studies on Steady
State and Dynamic Distortion,” Journal of Aircraft, Vol. 35, No. 4, pp. 513–552.
Anderson, J. D. Jr (2003), “The Wright Brothers Aerodynamics and the Future of Flight,”
AIAA Paper 2003-4299.
REFERENCES 445

Annaswamy, A. M., and Ghoniem, A. F. (2002), “Active Control of Combustion


Instability: Theory and Practice,” IEEE Control System Magazine, Vol. 22, No. 6, pp.
37–54.
Annaswamy, A. M., El-Rifai, O., Fleifilj, M., and Ghoniem, A. F. (1998), “A Model-Based
Self-Tuning Controller for Thermoacoustic Instability,” Combustion Science
Technology, Vol. 135, pp. 213–240.
Antoulas, A. C., Sorensen, D. C., and Gugercin, S. (2001), “A Survey of Model Reduc-
tion Methods for Large-Scale Systems,” Contemporary Mathematics, Vol. 280,
pp. 193–219.
Appukuttan, A., Shyy, W., Sheplak, M., and Cattafesta, L. (2003), “Mixed Convection
Induced by MEMS-Based Thermal Shear Stress Sensors,” Numerical Heat Transfer-A,
Vol. 43, No. 3, pp. 283–305.
Arad, E., Martin, P. B., Wilson, J., and Tung, C. (2006), “Control of Massive Separation on
a Thick-Airfoil Wing: A Computational and Experimental Study,” AIAA Paper
2006-0322.
Arakeri, V. H., Krothapalli, A., Siddavaram, V., Alkislar, M. B., and Lourenco, L. M.
(2003), “On the Use of Microjets to Suppress Turbulence in a Mach 0.9 Axisymmetric
Jet,” Journal of Fluid Mechanics, Vol. 490, pp. 75–98.
Arbey, H., and Ffowcs Williams, J. E. (1984), “Active Cancellation of Pure Tones in an
Excited Jet,” Journal of Fluid Mechanics, Vol. 149, pp. 445–454.
Armstrong, R. R., Michalke, A., and Fuchs, H. (1977), “Coherent Structures in Jet
Turbulence and Noise,” AIAA Journal, Vol. 15, No. 7, pp. 1011–1017.
Arndt, R. E. A., Long, D. F., and Glauser, M. N. (1997), “The Proper Orthogonal
Decomposition of Pressure Fluctuations Surrounding a Turbulent Jet,” Journal of Fluid
Mechanics, Vol. 340, pp. 1–33.
Arnold, D. P., Nishida, T., Cattafesta, L., and Sheplak, M. (2003), “MEMS-Based Acoustic
Array Technology,” Journal of the Acoustical Society of America, Vol. 113, No. 1,
pp. 289–298.
Arnoldi, W. (1951), “The Principle of Minimized Iterations in the Solution of the Matrix
Eigenvalue Problem,” Quarterly Applied Mathematics, Vol. 9, pp. 17–29.
Aronson, D. G., Betelu, S. I., and Kevrekidis, I. G. (2001), “Going with the Flow: a
Lagrangian Approach to Self-Similar Dynamics and its Consequences,” Preprint,
http://arXiv:nlin/0111055.
Arts, T., and Colton, T. (2004), “Investigation of a High Lift LP Turbine Blade submitted
to Passing Wakes: Part 2 Boundary Layer Transition,” AFRLT GT2004-53768.
Arunajatesan, S., Shipman, J. D., and Sinha, N. (2002), “Hybrid RANS-LES Simulation of
Cavity Flow Fields with Control,” AIAA Paper 2002-1130.
Arwatz, G., Fono, I., and Seifert, A. (2008), “Suction and Oscillatory Blowing Actuator,”
Proceedings of the IUTAM Symposium on Flow Control and MEMS, Royal Geographi-
cal Society, 19–22 Sept. 2006. Springer-Verlag, Berlin, pp. 33–44.
Ashcroft, G., and Schultz, J. (2004), “Numerical Modelling of Wake-jet Interaction
with Application to Active Noise Control in Turbomachinery,” AIAA Paper
2004-2655.
Åström, K. J., and Murray, R. M. (2008), Feedback Systems: An Introduction for Scientists
and Engineers, Princeton University Press, Princeton, NJ.
Attinello, J. S. (1961), “Design and Engineering Features of Flap Blowing Installations,”
Boundary Layer and Flow Control, edited by G. V. Lachmann, Vol. 1, Pergamon Press,
New York, pp. 463–515.
446 REFERENCES

Aubry, N., Holmes, P., Lumley, J. L., and Stone, E. (1988), “The Dynamics of Coherent
Structures in the Wall Region of a Turbulent Boundary Layer,” Journal of Fluid
Mechanics, Vol. 192, pp. 115–173.
Auer, M. P., Gebauer, K. G., Hirsch, C., Kmosl, C., and Sattelmayer, T. (2005), “Active
Instability Control: Feedback of Combustion Instabilities on the Injection of Gaseous
Fuel,” Journal of Engineering for Gas Turbines and Power, Vol. 127, pp. 749–754.
Ausseur, J., Pinier, J., and Glauser, M. (2006), “Flow Separation Control Using a Convection
Based POD Approach,” AIAA Paper 2006-3017.
Babinsky, H., and Ogawa, H. (2006), “Three-Dimensional SBLI Control for Transonic
Airfoils,” AIAA Paper 2006-3698.
Babinsky, H., Makinson, N., and Morgan, C. (2007), “Micro-Vortex Generator Flow
Control for Supersonic Engine Inlets,” AIAA Paper 2007-521.
Bachar, T. (2001), “Generating Dynamically Controllable Oscillatory Fluid Flow”,
US Patent 6,186,412, filed 13 Feb 2001.
Bake, S., Meyer, D. G. W., and Rist, U. (2002), “Turbulence Mechanism in Klebanoff
Transition: A Quantitative Comparison of Experiment and Direct Numerical
Simulation,” Journal of Fluid Mechanics, Vol. 459, pp. 217–243.
Baker, W. J., and Paterson, E. G. (2006), “Simulation of Steady Circulation Control for the
General Aviation Circulation Control (GACC) Wing,” Applications of Circulation
Control Technology, edited by R. D. Joslin, and G. S. Jones, Vol. 214, AIAA Progress
in Aeronautics and Astronautics Series, Reston, VA, pp. 513–537.
Balakumar, P. (2005), “Computations of Flow over a Hump Model Using Higher Order
Method with Turbulence Modeling,” AIAA Paper 2005-1270.
Balzer, W., Gross, A., and Fasel, H. F. (2007), “Active Flow Control of Low-Pressure
Turbine Separation,” Proceedings of the HPCMP Users Group Conference 2007, June
18–21, Pittsburgh, PA, Post, D. E. (Ed.), pp. 73–82.
Banaszuk, A., Jacobson, C. A., Khibnik, A. I., and Mehta, P. G. (1999), “Linear and
Nonlinear Analysis of Controlled Combustion Processes. Part I: Linear Analysis,”
Proceedings of the International Conference on Control Applications.
Banaszuk, A., Ariyur, K. B., Krstić, M., and Jacobson, C. A. (2004), “An Adaptive Algorithm
for Control of Combustion Instability,” Automatica, Vol. 40, pp. 1965–1972.
Banaszuk, A., Mehta, P. G., and Hagen, G. (2006a), “The Role of Control in Design: From
Fixing Problems to the Design of Dynamics,” Proceedings of International Symposium
on Advanced Control of Chemical Process, Gramado, Brazil.
Banaszuk, A., Mehta, P. G., Jacobson, C. A., and Khibnik, A. I. (2006b), “Limits of
Achievable Performance of Controlled Combustion Processes,” IEEE Transactions on
Control Systems Technology, Vol. 14, No. 5, pp. 881–895.
Barbosa, S., Garcia, M. de La Cruz, Ducruix, S., Labegorre, B., and Lacas, F. (2007),
“Control of Combustion Instabilities by Local Injection of Hydrogen,” Proceedings of
the Combustion Institute, Vol. 31, pp. 3207–3214.
Barkley, D., and Henderson, R. (1996), “Three-Dimensional Floquet Stability Analysis
of the Wake of a Circular Cylinder,” Journal of Fluid Mechanics, Vol. 322, pp.
215–241.
Barone, M. F., and Lele, S. K. (2005), “Receptivity of the Compressible Mixing Layer,”
Journal of Fluid Mechanics, Vol. 540, pp. 301–335.
Bar-Sever, A. (1989), “Separation Control on an Airfoil by Periodic Forcing,” AIAA
Journal, Vol. 27, No. 6, pp. 820–821.
REFERENCES 447

Baruzzini, D., Domel, N., and Miller, D. N. (2007), “Pulsed Injection Flow Control for
Throttling in Supersonic Nozzles — A Computational Fluid Dynamics Study,” AIAA
Paper 2007-4215.
Barwolff, G., Wengle, H., and Geggle, H. (1996), “Direct Numerical Simulation of
Transitional Backward-Facing Step Flow Manipulated by Oscillating Blowing/
Suction,” Engineering Turbulence Modelling and Experiments 3, edited by W. Rodi,
and G. Bergeles, Elsevier Science, Amsterdam, pp. 219–228.
Başar, T., and Bernhard, P. (1995), H•-Optimal Control and Related Minimax Design
Problems, Birkhäuser, Boston, MA.
Batchelor, G. K. (1964), “Axial Flow in Trailing Line Vortices,” Journal of Fluid Mechanics,
Vol. 20, Part 2, pp. 645–658.
Batten, P., Goldberg, U., and Chakravarthy, S. (2002), “LNS—an Approach towards
Embedded LES,” AIAA Paper 2002-0427.
Baurle, R. A., Tam, C.-J., Edwards, J. R., and Hassan, H. A. (2001), “An Assessment of
Boundary Treatment and Algorithm Issues on Hybrid RANS/LES Solution Strategies,”
AIAA Paper 2001-2562.
Bearman, P. W., and Currie, I. G. (1979), “Pressure-Fluctuation Measurements on
an Oscillating Circular Cylinder,” Journal of Fluid Mechanics, Vol. 91, Part 4,
pp. 661–677.
Bearman, P. W., and Harvey, J. K. (1976), “Golf Ball Aerodynamics,” The Aeronautical
Quarterly, Vol. 27, No. 2, pp. 112–122.
Bearman, P. W., and Harvey, J. K. (1993), “Control of Circular Cylinder Flow by the Use
of Dimples,” AIAA Journal, Vol. 31, No. 10, pp. 1753–1756.
Bearman, P. W., and Obasaju E. D. (1982), “An Experimental Study of Pressure Fluctuations
on Fixed and Oscillating Square-Section Cylinders,” Journal of Fluid Mechanics,
Vol. 110, pp. 207–321.
Bechert, D., and Pfizenmaier, E. (1975), “On the Amplification of Broadband Jet Noise by
a Pure Tone Excitation,” Journal of Sound and Vibration, Vol. 43, pp. 321–367.
Bechert, D. W., Bruse, M., Hage, W., van der Hoeven, J. G. T., and Hoppe, G. (1997),
“Experiments on Drag-Reducing Surfaces and Their Optimization with an Adjustable
Geometry,” Journal of Fluid Mechanics, Vol. 338, pp. 59–88.
Becker, J., and Hassa, C. (2003), “Liquid Fuel Placement and Mixing of Generic Aeroengine
Premix Module at Different Operating Conditions,” Journal of Engineering for Gas
Turbines and Power, Vol. 125, No. 4, pp. 901–908.
Becker, R., King, R., Petz, R., and Nitsche, W. (2006), “Adaptive Closed-Loop Separation
Control on a High-Lift Configuration Using Extremum Seeking,” AIAA Paper
2006-3493.
Behrouzi, P., and McGuirk, J. (2006), “Flow Control of Jet Mixing Using a Pulsed Fluid
Tab Nozzle,” AIAA Paper 2006-3509.
Bélanger, P. R. (1995), Control Engineering. A Modern Approach, Saunders College
Publishing Philadelphia, PA.
Bell, D. J., Lu, T. J., Fleck, N. A., and Spearing, S. M. (2005), “MEMS Actuators and
Sensors: Observations on their Performance and Selection for Purpose,” Journal of
Micromechanics and Microengineering, Vol. 15, pp. S153–S164.
Bellucci, V., Flohr, P., Magni, F., and Paschereit, C. O. (2004), “On the Use of Helmoholz
Resonators for Damping Acoustic Pulsation in Industrial Gas Turbines,” Journal of
Engineering for Gas Turbines and Power, Vol. 126, pp. 271–275.
448 REFERENCES

Ben-Hamou, E., Arad, E., and Seifert, A. (2007) “Generic Transport Aft-Body Drag Reduction
Using Active Flow Control,” Journal of Flow, Turbulence and Combustion—Special
Issue on Air-Jet Actuators and Their Use for Flow Control, Vol. 78, pp. 365–382.
Benard, N., Joliboius, Forte, M., Touchard, M., and Moreau, E. (2007), “Control of an
Axisymmetric Subsonic Air Jet by Plasma Actuator,” Experiments in Fluids, Vol. 43,
No. 4, pp. 603–616.
Bender, E., Miller, D., Smith, B., Yagle, P., Vermeulen, P., and Walker, S. (2000), “Simula-
tion of Pulsed Injection in a Crossflow Using 3-D Unsteady CFD,” AIAA Paper
2000-2318.
Béra, J.-C., Michard, M., Sunyach, M., and Comte-Bellot, G. (2000), “Changing Lift and
Drag by Jet Oscillation: Experiments on a Cylinder with Turbulent Separation,”
European Journal of Mechanics B. Fluid, Vol. 19, No. 5, pp. 575–595.
Berger, E. (1967), “Suppression of Vortex Shedding and Turbulence behind Oscillating
Cylinders,” Physics of Fluids, Vol. 10 (Suppl.), pp. 191–193.
Beris, A. N., and Dimitropoulos, C. D. (1999), “Pseudospectral Simulation of Turbulent
Viscoelastic Channel Flow,” Computer Methods in Applied Mechanics and Engineer-
ing, Vol. 180, pp. 365–392.
Beris, A. N., Dimitropoulos, C. D., Sureshkumar, R., and Handler, R. D. (2000), “Direct
Numerical Simulations of Polymer-Induced Drag Reduction in Viscoelastic Turbulent
Channel Flows,” XIIIth International Congress on Rheology, Cambridge, UK.
Bernal, L. P., and Roshko, A. (1986), “Streamwise Vortex Structure in Plane Mixing
Layers,” Journal of Fluid Mechanics, Vol. 170, pp. 499–525.
Bernhardt, J. E., and Williams, D. R. (2000), “Closed-Loop Control of Forebody Flow
Asymmetry,” Journal of Aircraft, Vol. 37, No. 3, pp. 491–498.
Bertlerud, A. (1998), “Transition on a Three-Element High Lift Configuration at High
Reynolds Numbers,” AIAA Paper 98-0703.
Bertolotti, F. P., Herbert, T., and Spalart, P. R. (1992), “Linear and Nonlinear Stability of
the Blasius Boundary Layer,” Journal of Fluid Mechanics, Vol. 242, pp. 441–474.
Bettini, C., and Cravero, C. (2007), “Computational Analysis of Flow Separation Control
for the Flow Over a Wall-Mounted Hump Using a Synthetic Jet,” AIAA Paper
2007-0516.
Betz, A. (1932), “Behavior of Vortex Systems,” Zeit. für angewandte Math. und Mech., Vol.
12, No. 3. (See NACA TM-713, June 1933).
Betz, A. (1961), “History of Boundary Layer Control in German,” Boundary Layer and
Flow Control, edited by G. V. Lachmann, Vol. 1, Pergamon Press, New York, pp. 1–20.
Bewley, T. R. (2001), “Flow Control: New Challenges for a New Renaissance,” Progress in
Aerospace Sciences, Vol. 37, No. 1, pp. 21–58.
Bewley, T. R., and Aamo, O. M. (2002), “On the Search for Fundamental Performance
Limitation in Fluid-Mechanical Systems,” ASME Paper FEDSM’02.
Bewley, T. R., and Liu, S. (1998), “Optimal and Robust Control and Estimation of Linear
Paths to Transition,” Journal of Fluid Mechanics, Vol. 365, pp. 305–349.
Bewley, T. R., Temam, R., and Ziane, M. (2000), “A General Framework for Robust Control
in Fluid Mechanics,” Physica D, Vol. 138, pp. 360–392.
Bewley, T. R., Moin, P., and Temam, R. (2001), “DNS-Based Predictive Control of
Turbulence: an Optimal Benchmark for Feedback Algorithms,” Journal of Fluid
Mechanics, Vol. 447, pp. 179–225.
Beyn, W.-J., and Thümmler, V. (2004), “Freezing Solutions of Equivariant Evolution
Equations,” SIAM Journal on Applied Dynamical Systems, Vol. 3, No. 2, pp. 85–116.
REFERENCES 449

Bhargava, A., Kendrick, D. W., Colket, M. B., Sowa, W. A., Casleton, K., and Maloney,
D. J. (2000), “Pressure Effects on NOx and CO Emission in Industrial Gas Turbines,”
ASME Paper no. 2000-GT-97.
Bhat, T. R. S., Anderson, B. A., and Gutmark, E. J. (2000), “Flexible Filaments in Jets and
the Interaction Mechanisms,” AIAA Paper 2000-0083.
Biedron, R., Vatsa, V., and Atkins, H. (2005), “Simulation of Unsteady Flows Using an
Unstructured Navier–Stokes Solver on Moving and Stationary Grids,” AIAA Paper
2005-5093.
Bilanin, A. J., and Widnall, S. E. (1973), “Aircraft Wake Dissipation by Sinusoidal
Instability and Vortex Breakdown,” AIAA Paper 73-107.
Billoud, G., Candel, S., Galland, M. A., and Huu, C. H. (1992), “Adaptive Active Control of
Combustion Instabilities,” Combustion Science and Technology, Vol. 81, pp. 257–283.
Bird, R. B., Armstrong, R. C., and Hassager, O. (1987), Dynamics of Polymeric Liquids,
John Wiley, New York.
Blackwelder, R. F., and Kaplan, R. E. (1972), “On the Wall Structure of the Turbulent
Boundary Layer,” Journal of Fluid Mechanics, Vol. 76, Issue 1, pp. 89–112.
Blake, W. K. (1986), Mechanics of Flow Induced Sound and Vibration, Academic Press,
New York.
Bloor, M. S. (1964), “The Transition to Turbulence in the Wake of a Circular Cylinder,”
Journal of Fluid Mechanics, Vol. 19, part 2, pp. 290–304.
Bloxham, M., Reimann, D., Crapo, K., Pluim, J., and Bons, J. P. (2007), “Synchronizing
Separation Flow Control with Unsteady Wakes in a Low-Pressure Turbine Cascade,”
2007 IGTI Conference, Montreal, Paper #GT2007-27529, May 2007.
Bloxsidge, G. J., Dowling, A. P., Hooper, N., and Langhorne, P. J. (1988), “Active Control
of Reheat Buzz,” AIAA Journal, Vol. 26, pp. 783–790.
Bonn, D., Amarouchène, Y., Wagner, C., Douady, S., and Cadot, O. (2005), “Turbulent
Drag Reduction by Polymers,” Journal of Physics: Condensed Matter, Vol. 17,
pp. S1195–1202.
Bonnet, J. P., Delville, J., Glauser, M. N., Antonia, R. A., Bisset, D. K., Cole, D. R., Fiedler,
H. E., Garem, J. H., Hilberg, D., Jeong, J., Kevlahan, N. K. R., Ukeiley, L. S., and
Vincendeau, E. (1998), “Collaborative Testing of Eddy Structure Identification Meth-
ods in Free Turbulent Shear Flows,” Experiments in Fluids, Vol. 25, pp. 197–225.
Bons, J. P., Sondergaard, R., and Rivir, R. B. (1999), “Control of Low-Pressure Turbine
Separation Using Vortex Generator Jets,” AIAA Paper 99-0367.
Bons, J. P., Sondergaard, R., and Rivir, R. B. (2001a), “Turbine Separation Control Using
Pulsed Vortex Generator Jets,” Journal of Turbomachinery, Vol. 123, No. 2,
pp. 198–206.
Bons, J. P., Taylor, R. P., McClain, S. T., and Rivir, R. B. (2001b), “The Many Faces of
Turbine Surface Roughness,” Journal of Turbomachinery, Vol. 123, No. 4,
pp. 739–748.
Bons, J. P., Sondergaard, R., and Rivir, R. B. (2002), “The Fluid Dynamics of LPT Blade
Separation Control Using Pulsed Jets,” Journal of Turbomachinery, Vol. 124, No. 1,
pp. 77–85.
Bons, J. P., Hansen, L. C., Clark, J. P., Koch, P. J., and Sondergaard, R. (2005), “Designing
Low-Pressure Turbine Blades with Integrated Flow Control,” Proceedings of GT2005,
ASME Turbo Expo 2005: Power for Land, Sea, and Air, June, Reno-Tahoe, NV.
Borgoltz, A., Craig, M., and Devenport, W. (2005), “Trailing Edge Blowing of Fan Blades,”
AIAA Paper 2005-3031.
450 REFERENCES

Borgoltz, A., Craig, M., and Devenport, W. (2006), “Space-Time Correlations and Trailing
Edge Flow Structure in Fan-Blade Wakes with Trailing Edge Blowing,” AIAA Paper
2006-2480.
Bousman, W. G. (1998), “A Qualitative Examination of Dynamic Stall from Flight Test
Data,” Journal of the American Helicopter Society, Vol. 43, pp. 279–295.
Bousman, W. G. (2000), “Evaluation of Airfoil Dynamic Stall Characteristics for
Maneuverability”, 26th European Rotorcraft Forum, The Hague, Netherlands,
pp. 38.1–38.21.
Boxwell, D. A., Schmitz, F. H., Splettstoeser, W. R., and Schultz, K. J. (1983), “Model
Helicopter Rotor High-Speed Impulsive Noise: Measured Acoustics and Blade
Pressures,” Ninth European Rotorcraft Forum, Stresa, Italy.
Boxx, I., Rivir, R., Newcamp, Lt J., Franke, M., and Woods, N. (2006a), “A PIV Study of
a Plasma Discharge Flow Control Actuator on a Flat Plate in an Aggressive Pressure
Induced Separation,” GT 2006-91044, IGTI, Barcelona.
Boxx, I., Rivir, R., Newcamp, J., and Woods, N. (2006b), “Reattachment of a Separated
Boundary Layer on a Flat Plate in a Highly Adverse Pressure Gradient Using a Plasma
Actuator,” AIAA Paper 2006-3023.
Bradbury, L. J. S., and Khadem, A. H. (1975), “The Distortion of a Jet by Tabs,” Journal of
Fluid Mechanics, Vol. 70, No. 4, pp. 801–813.
Brandt, T. (2007), “A Posteriori Study on Modelling and Numerical Error in LES Applying
the Smagorinsky Model,” Complex Effects in Large Eddy Simulations, Lecture Notes
in Computational Science and Engineering 56, edited by S. C. Kassinos, C. A. Langer,
G. Iaccarino, and P. Moin, Springer-Verlag, Berlin, pp. 171–189.
Brasseur, J. G., Robert, A., Collins, L. R., and Vaithianathan, T. (2005), “Fundamental
Physics Underlying Polymer Drag Reduction, from Homogenous DNS Turbulence
with the FENE-P Model,” 2nd International Symposium on Seawater Drag Reduction,
Busan, Korea.
Bray, K. N. C., Dowling, A. P., and Zhu, M. (2005), “Transfer Function Calculations for
Aeroengine Combustion Oscillations,” Journal of Engineering for Gas Turbines and
Power, Vol, 127, pp. 19–26.
Brear, M. J., Warfield, Z., Mangus, J. F., Braddom, S., Paduano, J. D., Philhower, J. S.
(2003), “Flow Separation within the Engine Inlet of an Uninhabited Combat Air
Vehicle (UCAV),” FEDSM 2003-45579.
Brehm, C., Gross, A., and Fasel, H. F. (2006), “Closed-Loop Control of Low-Pressure
Turbine Laminar Separation,” AIAA Paper 2006-3021.
Bridges, D. (2007), “Early Flight-Test and Other Boundary-Layer Research at Mississippi
State 1949–1960,” Journal of Aircraft, Vol. 44, No. 5, pp. 1635–1652.
Brion, V., Sipp, D., and Jacquin, L. (2007), “Optimal Amplification of the Crow Instability,”
Physics of Fluids, Vol. 19, p. 111703.
Bristol, R. L., Ortega, J. M., Marcus, P. S. and Savas, O. (2004), “On Cooperative
Instabilities of Parallel Vortex Pairs,” Journal of Fluid Mechanics, Vol. 517,
pp. 331–358.
Broadhurst, M., and Sherwin, S. J. (2008), “The Parabolized Stability Equations for
3D-Flows: Implementation and Numerical Stability,” Applied Numerical Mathematics,
Vol. 58, No. 7, pp. 1017–1029.
Broadhurst, M., Sherwin, S. J., and Theofilis, V. (2006), “Spectral Element Stability
Analysis of Vortical Flows,” IUTAM Laminar-Turbulent Symposium VI, edited by R.
Govindarajan, Bangalore, India, pp. 153–158.
REFERENCES 451

Brocklehurst, A., and Pike, A. C. (1994), “Reduction of BVI Noise Using a Vane Tip,” AHS
Aeromechanics Specialists Conference, San Francisco, CA.
Brookfield, J. M. (1997), “Turbofan Stator/Rotor Interaction Noise Reduction through
Trailing-edge blowing,” Ph.D. Thesis, MIT, Cambridge, MA.
Brookfield, J. M., and Waitz, I. A. (2000), “Trailing-Edge Blowing for Reduction of
Turbomachinery Fan Noise,” Journal of Propulsion and Power, Vol. 16, No. 1,
pp. 57–64.
Brooks, S. J., Cant, R. S., Dowling, A. P., and Dupere, I. D. J. (2001), “Computational
Modeling of Self-Excited Combustion Instabilities,” Journal of Engineering for Gas
Turbines and Power, Vol. 123, pp. 322–326.
Brooks, T. F. (1993), “Studies of Blade–Vortex Interaction Noise Reduction by Rotor Blade
Modification,” Noise-Con 93, May 1993, Williamsburg, VA.
Brooks, T. F., Booth, E. R., Jolly, R. J., Yeager, W. T., and Wilbur, M. L. (1989a), “Reduction
of Blade–Vortex Interaction Noise Using Higher Harmonic Pitch Control,” NASA TM
101624.
Brooks, T. F., Pope, D. S., and Marcolini, M. A. (1989b), “Airfoil Self Noise and Prediction,”
NASA-RP-1218.
Brooks, T. F., Booth, E. R., Jr., Boyd, D. D., Jr., Splettstoesser, W. R., Schultz, R.-J., Kubelf,
R., Niesl, G. H., and Strebytt, O. (1994), “Analysis of a Higher Harmonic Control Test
to Reduce Blade Vortex Interaction Noise,” Journal of Aircraft, Vol. 31, No. 6,
pp. 1341–1349.
Brown, C. A., and Bridges, J. (2006), “Acoustic Efficiency of Azimuthal Modes in Jet
Noise Using Chevron Nozzles,” AIAA Paper 2006-2645.
Brown, G. L., and Roshko, A. (1971), “The Effect of Density Differences on the Turbulent
Mixing Layer,” AGARD CPP-93, 23:1.
Brown, G. L., and Roshko, A. (1974), “On Density Effects and Large Structure in Turbulent
Mixing Layers,” Journal of Fluid Mechanics, Vol. 64, pp. 775–816.
Bruce, P. J. K., and Babinsky, H. (2007), “Micro-Vortex Generator Flow Control for
Supersonic Engine Inlets,” AIAA Paper 2007-0521.
Brücker, C., Spatz, J., and Schröder, W. (2005), “Wall Shear Stress Imaging using Micro-
Structured Surfaces with Flexible Micro-Pillars,” Experiments in Fluids, Vol. 39,
pp. 464–474.
Brüel and Kjær (1996), Microphone Handbook, Vol. 1: Theory, (BA51505) Copenhagen.
Burns, F. P. (1957), “Piezoresistive Semiconductor Microphone,” Journal of the Acoustical
Society of America, Vol. 29, No. 2, pp. 192–199.
Burns, J. A., and King, B. B. (1998), “A Reduced Basis Approach to the Design of
Low-Order Feedback Controllers for Nonlinear Continuous Systems,” Journal of
Vibration and Control, Vol. 4, No. 3, pp. 297–323.
Bush, R. H., and Mani, M. (2001), “A Two-Equation Large Eddy Stress Model for High
Sub-Grid Shear,” AIAA Paper 2001-2561.
Butler, K. M., and Farrell, B. F. (1992), “Three-Dimensional Optimal Perturbations in
Viscous Flows,” Physics of Fluids A, Vol. 4, No. 8, pp. 1637–1650.
Cain, A. B., Kerschen, E. J., Tassy, J. M., and Raman, G. (2004), “Simulations of Powered
Resonance Tubes: Helmholtz Resonator Geometries,” AIAA Paper 2004-2690.
Calkins, F. T., Butler, G. W., and Mabe, J. H. (2006), “Variable Geometry Chevrons for Jet
Noise Reduction,” AIAA Paper 2006-2546.
Callender, B., Gutmark, E., and Martens, S. (2005), “Far-field Acoustic Investigation into
Chevron Nozzle Mechanisms and Trends,” AIAA Journal, Vol. 43, No. 1, pp. 87–95.
452 REFERENCES

Candel, S. (1992), “Combustion Instabilities Coupled by Pressure Waves and their Active
Control,” Proceedings of the Combustion Institute, Vol. 24, pp. 1277–1296.
Candel, S. (2002), “Combustion Dynamics and Control: Progress and Challenges,”
Proceedings of the Combustion Institute, Vol. 29, pp. 1–28.
Candel, S., and Poinsot, W. L. (1987), “Active Control of Combustion Instability,”
Combustion and Flame, Vol. 70, pp. 281–289.
Candel, S., Poinsot, T., Samaniego, J. M., and Yip, B. (1993), “Low-Frequency Combustion
Instability Mechanisms in a Side-Dump Combustor,” Combustion and Flame, Vol. 94,
pp. 363–370.
Capizzano, F., Catalano, P., Marongiu, C., and Vitagliano, P. L. (2005), “U-RANS
Modelling of Turbulent Flows Controlled by Synthetic Jets,” AIAA Paper
2005-5015.
Caradonna, F. X., and Tung, C. (1981), “Experimental and Analytical Studies of a Model
Helicopter in Hover,” NASA TM 81232.
Caradonna, F. X., Lautenschlager, J. L., and Silva, M. J. (1988), “An Experimental Study
of Rotor–Vortex Interactions,” AIAA Paper 88-0045.
Carlin, G., Dadone, L., and Spencer, R. (1989), “Results of an Experimental Investigation
of Blade Tip Vortex Modification Devices,” NASA CR-181853.
Carmichael, B. H. (1981), “Low Reynolds Number Airfoil Survey,” NASA Contractor
Report 165803, Vol. I.
Carpenter, P. W., and Green, P. N. (1997), “The Aeroacoustics and Aerodynamics of High-
Speed Coanda Devices, Part 1: Conventional Arrangement of Exit Nozzle and Surface,”
Journal of Sound and Vibration, Vol. 208, No. 5, pp. 777–801.
Carpenter, P. W., and Smith, C. (1997), “The Aeroacoustics and Aerodynamics of High-
Speed Coanda Devices, Part 2: Effects of Modifications for Flow Control and Noise
Reduction,” Journal of Sound and Vibration, Vol. 208, No. 5, pp. 803–822.
Carpy, S., and Manceau, R. (2006), “Turbulence Modelling of Statistically Periodic Flows:
Synthetic Jet into Quiescent Air,” International Journal of Heat and Fluid Flow,
Vol. 27, No. 5, pp. 756–767.
Carr, L. W. (1988), “Progress in the Analysis and Prediction of Dynamic Stall,” Journal of
Aircraft, Vol. 25, No. 1, pp. 6–17.
Carr, L. W., and Chandrasekhara, M. S. (1996), “Compressibility Effects on Dynamic
Stall,” Progress in Aerospace Sciences, Vol. 32, pp. 523–573.
Carr, L. W., and McAlister, K. W. (1983), “The Effect of a Leading-Edge Slat on the
Dynamic Stall of an Oscillating Airfoil,” AIAA Paper 83-2533.
Casey, J. P. (2004), “Effect of Dimple Pattern on the Suppression of Boundary Layer Separa-
tion on a Low Pressure Turbine Blade,” AFIT MS Thesis, AFIT/GAE/ENY/04-M05.
Catalano, P., Wang, M., Iaccarino, G., Sbalzarini, I. F., and Moumoutsakos, P. (2002),
“Optimization of Cylinder Flow Control via Actuators with Zero Net Mass Flux,”
Proceedings of the Summer Program 2002, Center for Turbulence Research, NASA
Ames/Stanford Univ., pp. 297–303.
Cathalifaud, P., and Luchini, P. (2000), “Algebraic Growth in Boundary Layers: Optimal
Control by Blowing and Suction,” European Journal of Mechanics B/Fluids, Vol. 19,
pp. 469–490.
Cattafesta, L. N., III, Garg, S., Choudhari, M., and Li, F. (1997), “Active Control of Flow-
Induced Cavity Resonance,” AIAA Paper 97-1804.
Cattafesta, L. N., III, Garg, S., and Shukla, D. (2001), “The Development of Piezoelectric
Actuators for Active Flow Control,” AIAA Journal, Vol. 39, No. 8, pp. 1562–1568.
REFERENCES 453

Cattafesta, L. N., III, Williams, D. R., Rowley, C. W., and Alvi, F. (2003), “Review of
Active Control of Flow-Induced Cavity Resonance,” AIAA Paper 2003-3567.
Chan, Y. Y. (1974), “Spatial Waves in Turbulent Jets,” Physics of Fluids, Vol. 17, No. 1,
pp. 46–53.
Chang, P. K. (1961), “Drag Reduction of an Airfoil by Injection of Sound Energy,” Journal
of Aeronautical Science, Vol. 28, No. 9, pp. 742–743.
Chang, R. C., Hsiao, F. B., and Shyu, R. N. (1992), “Forcing Level Effects of Internal
Acoustic Excitation on the Improvement of Airfoil Performance,” AIAA Journal,
Vol. 29, No. 5, pp. 823–829.
Chang, Y., Collis, S. S., and Ramakrishnan, S. (2002), “Viscous Effects in Control of Near-
Wall Turbulence,” Physics of Fluids, Vol. 14, No. 11, pp. 4069–4080.
Chang, P. A., III, Slomski, J., Marino, T., and Ebert, M. P. (2005), “Numerical Simulation
of Two- and Three-Dimensional Circulation Control Problems,” AIAA Paper
2005-0080.
Chang, P. A., III, Slomski, J., Marino, T., Ebert, M. P., and Abramson, J. (2006), “Full
Reynolds-Stress Modeling of Circulation Control Airfoils,” Applications of Circulation
Control Technology, edited by R. D. Joslin, and G. S. Jones, Vol. 214, AIAA Progress
in Aeronautics and Astronautics Series, Reston, VA, pp. 445–466.
Chapman, D. R. (1979), “Computational Aerodynamics Development and Outlook,” AIAA
Journal, Vol. 17, No. 12, pp. 1293–1313.
Charles, B. D., Tadghighi, H., and Hassan, A. A. (1996), “Higher Harmonic Actuation of
Trailing-Edge Flaps for Rotor BVI Noise Control,” 52nd Annual Forum of the American
Helicopter Society, Washington, DC.
Chase, A. (1981), “A Slice of Golf,” Science, Vol. 81, No. 2, pp. 90–91.
Chatterjee, A. (2000), “An Introduction to the Proper Orthogonal Decomposition,” Current
Science, Vol. 78, No. 7, pp. 808–817.
Chen, Y., Liang, K., Aung, A., Glezer, A., and Jagoda, J. (1999), “Enhanced Mixing in a
Simulated Combustor Using Synthetic Jet Actuators,” AIAA Paper 1999-0449.
Chen, F.-J., Yao, C., Beeler, G. B., Bryant, R. G., and Fox, R. L. (2000), “Development of
Synthetic Jet Actuators for Active Flow Control at NASA Langley,” AIAA Paper
2000-2405.
Chevalier H. (1973), “Flight Test Studies of the Formation and Dissipation of Trailing
Vortices,” Journal of Aircraft, Vol. 10, No. 1, pp. 14–18.
Chevalier, M., Hoegberg, M., Berggren, M., and Henningson, D. S. (2002), “Linear and
Nonlinear Optimal Control in Spatial Boundary Layers,” AIAA Paper 2002-2755.
Chevalier, M., Hoepffner, J., Akervik, E., and Henningson, D. S. (2007), “Linear Feedback
Control and Estimation Applied to Instabilities in Spatially Developing Boundary
Layers,” Journal of Fluid Mechanics, Vol. 588, pp. 163–187.
Chiarelli, C., Johnsen, R. K., Shieh, C. F., and Wing, D. J. (1993), “Fluidic Scale Model
Multi-Plane Thrust Vector Control Test Results,” AIAA Paper 93-2433.
Cho, Y., Chopra, J., and Morris, P. J. (2007), “Immersed Boundary Method for Compress-
ible High-Reynolds Number Viscous Flow around Moving Bodies,” AIAA Paper
2007-0125.
Choi, H. (1993), “Toward Large Eddy Simulation of Turbulent Flow over an Airfoil,”
Annual Research Briefs, Center for Turbulence Research, NASA Ames/Stanford Univ.,
pp. 145–149.
Choi, H., Moin, P., and Kim, J. (1994), “Active Turbulence Control for Drag Reduction in
Wall-Bounded Flow,” Journal of Fluid Mechanics, Vol. 262, pp. 75–110.
454 REFERENCES

Choi, H., Hinze, M., and Kunisch K. (1999), “Instantaneous Control of Backward-Facing
Step Flows,” Applied Numerical Mathematics, Vol. 31, pp. 133–158.
Choi S., Choi H., and Kang S. (2002), “Characteristics of Flow over a Rotationally
Oscillating Cylinder at Low Reynolds Number,” Physics of Fluids, Vol. 14, No. 8,
pp. 2767–2776.
Choi, J. J., Annaswamy, A. M., Lou, H., and Alvi, F. S. (2006), “Active Control of Supersonic
Impingement Tones Using Steady and Pulsed Microjets,” Experiments in Fluids,
Vol. 41, No. 6, pp. 841–855.
Choi, H., Jeon, W.-P., and Kim, J. (2008), “Control of Flow over a Bluff Body,” Annual
Review of Fluid Mechanics, Vol. 40, pp. 113–139.
Chu, B. T., and Kovasznay, L. S. G. (1958), “Non-Linear Interactions in a Viscous Heat-
Conducting Compressible Gas,” Journal of Fluid Mechanics, Vol. 3, pp. 494–514.
Cierpka, C., Weier, T., and Gerbeth, G. (2007), “Electromagnetic Control of Separated
Flows using Periodic Excitation with Different Wave Forms,” Conference on Active
Flow Control, Sept. 2006, Berlin.
Citriniti, J. H., and George, W. K. (2000), “Reconstruction of the Global Velocity Field in
the Axisymmetric Mixing Layer Utilizing the Proper Orthogonal Decomposition,”
Journal of Fluid Mechanics, Vol. 418, pp. 137–166.
Cizmas, P., and Dorney, D. (1998), “Parallel Computation of Turbine Blade Clocking,”
AIAA Paper 98-3598.
Cizmas, P., and Miller, A. (2000), “Numerical Simulation of Flow on Turbine Blades with
Dimples,” Texas A & M Univ. AFRL Summer Faculty Report.
Clark, R. L. (1979), “Evaluation of F-111 Weapon Bay Aero-Acoustic and Weapon
Separation Improvement Techniques,” AFFDL TR-79-3003.
Claus, R. W., and Vanka, S. P. (1992), “Multigrid Calculations of a Jet in Crossflow,”
Journal of Propulsion and Power, Vol. 8, No. 2, pp. 425–431.
Coe, D. J., Allen, M. G., Trautman, M. A., and Glezer, A. (1994), “Micromachined Jets for
Manipulation of Macro Flows,” Solid-State Sensor and Actuator Workshop, Hilton
Head, SC, pp. 243–247.
Cohen, J. (1985), “Instabilities and Resonances in Turbulent Free Shear Flows,” Ph.D.
Thesis, Univ. Arizona, Tucson, AZ.
Cohen, J. M., and Banaszuk, A. (2003), “Factors Affecting the Control of Unstable
Combustors,” Journal of Propulsion and Power, Vol. 19, No. 5, pp. 811–821.
Cohen, J. M., and Banaszuk, A. (2005), “Factors Affecting the Control of Unstable
Combustors,” Combustion Instabilities in Gas Turbine Engines, edited by T. Lieuwen,
and V.Yang, Vol. 210, Progress in Astronautics and Aeronautics, pp. 581–610.
Cohen, J. M., and Rey, N. M. (1999), “Active Control of Combustion Instability in a
Liquid-Fueled Low-NOx Combustor,” Journal of Engineering for Gas Turbines and
Power, Vol. 121, p. 281–285.
Cohen J., and Wygnanski, I. (1987a), “The Evolution of Instabilities in the Axisymmetric
Jet. Part 1. The Linear Growth of Disturbances near the Nozzle,” Journal of Fluid
Mechanics, Vol. 176, pp. 191–219.
Cohen J., and Wygnanski, I. (1987b), “The Evolution of Instabilities in the Axisymmetric
Jet. Part 2. The Flow Resulting from the Interaction between Two Waves,” Journal of
Fluid Mechanics, Vol. 176, pp. 221–235.
Cohen, J. M., Proscia, W., and Stufflebeam, J. H. (2001), “The Effect of Fuel/Air Mixing
on Actuation Authority in an Active Combustion Instability Control System,” Journal
of Engineering for Gas Turbines and Power, Vol. 123, pp. 537–542.
REFERENCES 455

Coker, A., Neumeier, Y., Zinn, B. T., Menon, S., and Lieuwen, T. (2006), “Active Instability
Control Effectiveness in a Liquid Fueled Combustor,” Combustion Science and
Technology, Vol., 178, pp. 1251–1261.
Colby, J. A., Menon, S., and Jagoda, J. (2006), “Spray and Emission Characteristics near
Lean Blowout in a Counter-Swirl Stabilized Gas Turbine Combustor,” ASME Turbo
Expo 2006, Barcelona, Spain, GT2006-90974.
Cole, H. A., Jr., and Holleman, E. C. (1958), “Measured and Predicted Dynamic Response
Characteristics of a Flexible Airplane to Elevator Control over a Frequency Range
Including Three Structural Modes,” NACA TN-4147.
Colinet, P., Legros, J. C., and Velarde, M. G. (2001), Nonlinear Dynamics of Surface-
Tension-Driven Instabilities, Wiley-VCH, Berlin.
Collins, F. G., and Zelenevitz, J. (1975), “Influence of Sound upon Separated Flow over
Wings,” AIAA Journal, Vol. 13, No. 3, pp. 408–410.
Collis, S. S., and Lele, S. K. (1999), “Receptivity to Surface Roughness Near a Swept
Leading Edge,” Journal of Fluid Mechanics, Vol. 380, pp. 141–168.
Collis, S. S., and Dobrinsky, A. (2000), “Evaluation of Adjoint-Based Methods for the
Prediction of Receptivity,” IUTAM Laminar-Turbulent Symposium V, edited by W.
Saric and H. Fasel, Sedona, AZ, pp. 111–116.
Collis, S. S., Ghayour, K., Heinkenschloss, M., Ulbrich, M., and Ulbrich, S. (2001),
“Numerical Solution of Optimal Control Problems Governed by the Compressible
Navier–Stokes Equations,” International Series of Numerical Mathematics, Vol. 139,
Birkhäuser Verlag, Basel, pp. 43–55.
Collis, S. S., Chang, Y., Kellogg, S., and Prabhu, R. D. (2000), “Large Eddy Simulation and
Turbulence Control,” AIAA Paper 2000-2564.
Collis, S. S., Ghayour, K., Heinkenschloss, M., Ulbrich, M., and Ulbrich, S. (2002),
“Optimal Control of Unsteady Compressible Viscous Flows,” International Journal of
Numerical Methods in Fluids, Vol. 40, pp. 1401–1429.
Collis, S. S., Joslin, R. D., Seifert, A., and Theofilis, V. (2004), “Issues in Active Flow
Control: Theory, Control, Simulation, and Experiment,” Progress in Aerospace
Sciences, Vol. 40, No. 4–5, pp. 237–289.
Colonius, T. (2001), “An Overview of Simulation, Modeling, and Active Control of Flow/
Acoustic Resonance in Open Cavities,” AIAA Paper 2001-0076.
Compton, D. A., and Johnston, J. P. (1992), “Streamwise Vortex Production by Pitched
and Skewed Jets in a Turbulent Boundary Layer,” AIAA Journal, Vol. 30, No. 3,
pp. 640–647.
Conrad, T., Bibik, A., Shcherbik, D., Lubarsky, E., and Zinn, B. T. (2004), “Control of
Instabilities in Liquid Fueled Combustor by Modification of the Reaction Zone using
Smart Fuel Injector,” AIAA Paper 2004-4029.
Corbett, P., and Bottaro, A. (2001), “Optimal Linear Growth in Swept Boundary Layers,”
Journal of Fluid Mechanics, Vol. 435, pp. 1–23.
Corcos, G. M. (1959), “Some Effects of Sound Reduction Devices on a Turbulent Jet,”
Journal of Aero/Space Sciences, Vol. 26, pp. 717–722.
Corcos, G. M. (1963), “Resolution of Pressure in Turbulence,” Journal of the Acoustical
Society of America, Vol. 35, No. 2, pp. 192–199.
Corke, C., and Post, M. L. (2005), “Overview of Plasma Flow Control: Concepts,
Optimization, and Applications,” AIAA Paper 2005-0563.
Corke, T. C., He, C., and Patel, M. P. (2004), “Plasma Flaps and Slats: An Application of
Weakly Ionized Plasma Actuators,” AIAA Paper 2004-2127.
456 REFERENCES

Corke, T. C, Post, M. L., and Orlov, D. M. (2007), “SDBD Plasma Enhanced Aerodynamics:
Concepts, Optimization and Applications,” Progress in Aerospace Sciences, Vol. 43,
pp. 193–217.
Cornish, J. J., III (1953), “Prevention of Turbulent Separation by Suction through a
Perforated Surface,” Aerophysics Dept., Mississippi State Univ., Research Report
No. 7.
Corrsin, S., and Kistler, A. L. (1955), “Free-Stream Boundaries of Turbulent Flows,”
NACA TR 1244, Washington, DC.
Cortelezzi, L., Lee, K. H., Kim, J., and Speyer, J. L. (1998), “Skin-Friction Drag Reduction
via Robust Reduced-order Linear Feedback Control,” International Journal of
Computational Fluid Dynamics, Vol. 11, No. 1–2, pp. 79–92.
Cossu, C., and Brandt, L. (2002), “Stabilization of Tollmien-Schlicting Waves by Finite-
Amplitude Optimal Streaks in the Blasius Boundary Layer,” Physics of Fluids, Vol. 14,
No. 8, pp. L57–L60.
Cox, D. W. Jr., and Marble, F. E. (1953), “Servo-Stabilization of Low-Frequency
Oscillations in a Liquid Bipropellant Rocket Motor,” American Rocket Society Journal,
Vol. 23, pp. 63–69.
Crichton, D., de la Rosa Blanco, E., and Law, T. R. (2007), “Design and Operation for Ultra
Low-Noise Take-off,” AIAA Paper 2007-0456.
Crighton, D., and Ffowcs Williams, J. E. (1969), “Sound Generation by Turbulent
Two-Phase Flow,” Journal of Fluid Mechanics, Vol. 36, pp. 585–603.
Criminale, W. O., Jackson, T. L., and Joslin, R. D. (2003), Theory and Computation of
Hydrodynamic Stability, Cambridge University Press, Cambridge, UK.
Crittenden, T. M., Glezer, A., Funk, R., and Parekh, D. (2001), “Combustion-Driven Jet
Actuators for Flow Control,” AIAA Paper 2001-2768.
Crittenden, T. M., Shlyubsky, D., and Glezer, A. (2004), “Combustion-Driven Jet Actuators
in Reversed Flow Configurations,” AIAA Paper 2004-2689.
Crocco, L., Garrje, D. T., and Grey, J. (1960), “Theory of Liquid Propellant Rocket
Combustion Instability and Its Experiment Validation,” American Rocket Society
Journal, Vol. 30, pp. 159–168.
Crouch, J. D. (1992), “Localized Receptivity of Boundary Layers,” Physics of Fluids,
Vol. 4, p. 1408.
Crouch J. D. (1997), “Instability and Transient Growth for Two Trailing-Vortex Pairs,”
Journal of Fluid Mechanics, Vol. 350, pp. 311–330.
Crouch, J. D., Miller, G. D., and Spalart, P. R. (2001), “Active-Control System for Breakup
of Airplane Trailing Vortices,” AIAA Journal, Vol. 39, No. 12, pp. 2374–2381.
Crow, S. C. (1970), “Stability Theory for a Pair of Vortices,” AIAA Journal, Vol. 8,
pp. 2172–2179.
Crow, S. C., and Bate E. R. (1976), “Lifespan of Trailing Vortices in a Turbulent
Atmosphere,” Journal of Aircraft, Vol. 13, No. 7, pp. 476–482.
Crow, S. C., and Champagne, F. H., (1971), “Orderly Structures in Jet Turbulence,” Journal
of Fluid Mechanics, Vol. 48, pp. 491–547.
Cui, J., and Agarwal, R. K. (2005), “3-D CFD Validation of an Axisymmetric Jet in Cross-
Flow,” AIAA Paper 2005-1112.
Cui, J., and Agarwal, R. K. (2006), “Three-Dimensional Computation of a Synthetic Jet in
Quiescent Air,” AIAA Journal, Vol. 44, No. 12, pp. 2857–2865.
Culick, F. E. C., and Yang, V. (1995), “Overview of Combustion Instabilities in Liquid
Propellant Rocket Engines,” Liquid Rocket Engine Combustion Instability, edited by
REFERENCES 457

V. Yang, and W. E. Anderson, Vol. 169, Progress in Astronautics and Aeronautics,


pp. 3–37.
Curle, N. (1955), “The Influence of Solid Boundaries upon Aerodynamic Sound,”
Proceedings of the Royal Society of London, Vol. A231 pp. 505–514.
Currier, J. M., and Fung, K.-Y. (1992), “Analysis of the Onset of Dynamic Stall,” AIAA
Journal, Vol. 30, No. 10, pp. 2469–2477.
Curtain, R. F., and Zwart, H. J. (1995), An Introduction to Infinite-Dimensional Linear
System Theory, Springer-Verlag, New York.
Curtis, E., Hodson, H., Banieghbal, M., Denton, J., Howell, R., and Harvey, N. (1997),
“Development of Blade Profiles for Low Pressure Turbine Applications,” ASME
Journal of Turbomachinery, Vol. 119, No. 3, pp. 531–538.
Cutler, A. D., Beck, B. T., Wilkes, J. A., Drummond, P. J., Alderfer, W. D., Paul, M., and
Danehy, M. P. (2005), “Development of a Pulsed Combustion Actuator for High-Speed
Flow Control,” AIAA Paper 2005-1084.
Cybyk, B., Grossman, K., and Van Wie, D. (2003), “Computational Assessment of the
SparkJet Flow Control Actuator,” AIAA Paper 2003-3711.
Cybyk, B., Grossman, K., and Wilkerson, J. (2004), “Performance Characteristics of the
SparkJet Flow Control Actuator,” AIAA Paper 2004-2131.
Cybyk, B., Grossman, K., and Wilkerson, J. (2005), “Single-Pulse Performance of the
SparkJet Flow Control Actuator,” AIAA Paper 2005-0401.
Dahai, G., Cary, A. W., and Agarwal, R. K. (2003), “Numerical Simulation of Vectoring
of a Primary Jet with a Synthetic Jet,” AIAA Journal, Vol. 41, No. 12, pp. 2364–2370.
Dalley, S., and Oleson, J. P. (2003), “Sennacherib, Archimedes, and the Water Screw:
The Context of Invention in the Ancient World,” Technology and Culture, Vol. 44,
No. 1.
Damevin, H.-M., and Hoffman, K. A. (2002), “Numerical Simulations of Magnetic Flow
Control in Hypersonic Chemically Reacting Flows,” Journal of Thermophysics and
Heat Transfer, Vol. 16, No. 4, pp. 498–507.
Dandois, J., Garnier, E., and Sagaut, P. (2006a), “Unsteady Simulation of Synthetic Jet in
a Crossflow,” AIAA Journal, Vol. 44, No. 2, pp. 225–238.
Dandois, J., Garnier, E., and Sagaut, P. (2006b), “DNS/LES of Active Separation Control
by Synthetic Jets,” AIAA Paper 2006-3026.
Dannenberg, R. E., and Weiberg, J. A. (1942), “Section Characteristics of a 10.5-Percent-
Thick Airfoil with Area Suction as Affected by Chordwise Distribution of Permeability,”
NACA TN-2847.
Darabi, A., and Wygnanski, I. (2004a), “Active Management of Naturally Separated Flow
over a Solid Surface. Part 1. The Forced Reattachment Process,” Journal of Fluid
Mechanics, Vol. 510, pp. 105–129.
Darabi, A., and Wygnanski, I. (2004b), “Active Management of Naturally Separated Flow
over a Solid Surface. Part 2. The Separation Process,” Journal of Fluid Mechanics,
Vol. 510, pp. 131–144.
Datta, B. N. (2004), Numerical Methods for Linear Control Systems. Elsevier, San
Diego, CA.
Dausman, G. (1990), “The LTV XC-142 Experimental Aircraft Lessons Learned,” AIAA
Paper 1990-3204.
Davies, M. E. (1976), “A Comparison of the Wake Structure of a Stationary and Oscillating
Bluff Body, using a Conditional Averaging Technique”, Journal of Fluid Mechanics,
Vol. 75, Part 2, pp. 209–231.
458 REFERENCES

Dawson, S., Marcolini, M., Booth, E., Straub, F., Hassan, A. A., Tadghighi, H., and Kelly,
H. (1995), “Wind Tunnel Test of an Active Flap Rotor: BVI Noise and Vibration
Reduction,” 51st Forum of the American Helicopter Society, May.
De Angelis, E., Casiola, C. M., and Piva, R. (2002), “DNS of Wall Turbulence: Dilute
Polymers and Self-Sustaining Mechanisms,” Computers and Fluids, Vol. 31, pp. 495–507.
De la Rosa Blanco, E., Hall, C. A., and Crichton, D. (2007), “Challenges in the Silent
Aircraft Engine Design,” AIAA Paper 2007-0454.
De Palma, P., De Tullio, M. D., Pascazio, G., and Napolitano, M. (2006), “An Immersed-
Boundary Method for Compressible Viscous Flows,” Journal of Computers & Fluids,
Vol. 35, pp. 693–702.
Decher, R., and Mayer, D. W. (1994), “On Supersonic Inlet/Engine Stability,” AIAA Paper
94-3371.
Deere, K. A. (2000), “Computational Investigation of the Aerodynamic Effects on Fluidic
Thrust Vectoring,” AIAA Paper 2000-3598.
Deere, K. A. (2003), “Summary of Fluidic Thrust Vectoring Research Conducted at NASA
Langley Research Center,” AIAA Paper 2003-3800.
Deere, K. A., and Wing, D. J. (1998), “PAB3D Simulations of a Nozzle with Fluidic
Injection for Yaw-Thrust-Vector Control,” AIAA Paper 98-3254.
Deere, K. A., Berrier, B. L., Flamm, J. D., and Johnson, S. K. (2003), “Computational
Study of Fluidic Thrust Vectoring Using Separation Control in a Nozzle,” AIAA Paper
2003-3803.
Deere, K. A., Berrier, B. L., Flamm, J. D., and Johnson, S. K. (2005), “A Computational
Study of a Dual Throat Fluidic Thrust Vectoring Nozzle Concept,” AIAA Paper 2005-
3502.
Deere, K. A., Flamm, J. D., Berrier, B. L., and Johnson, S. K. (2007), “Computational
Study of an Axisymmetric Dual Throat Fluidic Thrust Vectoring Nozzle Concept for a
Supersonic Business Jet,” AIAA Paper 2007-5085.
Dejoan, A., and Leschziner, M. A. (2004), “Large Eddy Simulation of Periodically
Perturbed Separated Flow Over a Backward-Facing Step,” International Journal of
Heat and Fluid Flow, Vol. 25, pp. 581–592.
Dejoan, A., and Leschziner, M. A. (2005), “Large Eddy Simulation of a Plane Turbulent
Wall Jet,” Physics of Fluids, Vol. 17, 25–41.
DeLaat, J. C., Breisacher, K. J., Saus, J. R., and Paxson, D. E. (2000), “Active Combustion
Control for Aircraft Gas Turbine Engines,” NASA TM-2000-210346, (also AIAA
Paper 2000-3500).
DeMeis, R. (1988), “Stick-to-it Riblets,” Aerospace America, pp. 48–49.
Demuren, A. O. (1993), “Characteristics of Three-Dimensional Turbulent Jets in Cross-
flow,” International Journal of Engineering Science, Vol. 31, No. 6, pp. 899–913.
Deng, S., Jiang, L., and Liu, C. (2007), “DNS for Flow Separation Control Around an
Airfoil by Pulsed Jets,” Computers & Fluids, Vol. 36, No. 6, pp. 1040–1060.
Denton, J. (1993), “Loss Mechanisms in Turbomachines,” Journal of Turbomachinery,
Vol. 115, No. 4, pp. 621–656.
DeSalvo, M., and Glezer, A. (2006), “Aerodynamic Control at Low Angles of Attack using
Trapped Vorticity Concentrations,” AIAA Paper 2006-0100.
Dimitropoulos, C. D., Sureshkumar, R., and Beris, A. N. (1998), “Direct Numerical
Simulation of Viscoelastic Turbulent Channel Flow Exhibiting Drag Reduction: Effect
of the Variation of Rheological Parameters,” Journal of Non-Newtonian Fluid
Mechanics, Vol. 79, pp. 443–468.
REFERENCES 459

Dindar, M., Jansen, K., and Hassan, A. A. (1999), “Effects of Transpiration Flow Control
on Hovering Rotor Blades,” AIAA Paper 99-3192.
Dobrinsky, A. (2002), “Adjoint Analysis for Receptivity Prediction,” PhD. Thesis, Rice
University, Houston, TX.
Dobrinsky, A., and Collis, S. S. (2000), “Adjoint Parabolized Stability Equations for
Receptivity Prediction,” AIAA Paper 2000-2651.
Docquier, N., and Candel, S. (2002), “Combustion Control and Sensors: A Review,”
Progress in Energy and Combustion Science, Vol. 28, pp. 107–150.
Donald, D. (ed.) (1997), The Complete Encyclopedia of World Aircraft, Barnes & Noble,
New York.
Donaldson, C. duP., Snedeker, R. S., and Sullivan, R. D. (1974), “Calculation of Aircraft
Wake Velocity Profiles and Comparison with Experimental Measurements,” Journal of
Aircraft, Vol. 11, No. 9, pp. 547–555.
Dorato, P., Abdallah, C., and Cerone, V. (2000), Linear Quadratic Control, Krieger
Publishing, Malabar, FLE.
Dorney, D. J., and Davis, R. L. (1992), “Navier–Stokes Analysis of Turbine Blade Heat
Transfer and Performance,” ASME Journal of Turbomachinery, Vol. 114, No. 4,
pp. 795–806.
Dorney, D. J., and Sharma, O. P. (1996), “A Study of Turbine Performance Increases
Through Airfoil Clocking,” AIAA Paper 96-2816.
Dorney, D. J., Sharma, O. P., and Gundy-Burlet, K. L. (1998), “Physics of Airfoil Clocking
in a High-Speed Axial Compressor,” ASME Paper 98-GT-082.
Dowell, E. H. (1980), “Nonlinear Aeroelasticity,” New Approaches to Nonlinear Problems
in Dynamics, edited by P. J. Holmes, SIAM Publications, Philadelphia, PA,
pp. 147–172.
Dowling, A. P. (2000), “Vortex, Sound and Flames—Damaging Combination,” Aeronautical
Journal, Vol. 104, pp. 105–116.
Dowling, A. P., and Dupere, I. D. J. (2005), “The Use of Helmholtz Resonators in a Prac-
tical Combustor,” Journal of Engineering for Gas Turbines and Power, Vol. 127,
pp. 268–274.
Dowling, A. P., and Ffowcs Williams, J. E., (1983), Sound and Sources of Sound, Ellis
Horwood, Chichester, UK.
Dowling, A. P., and Morgans, A. S. (2005), “Feedback Control of Combustion Oscillations,”
Annual Review of Fluid Mechanics, Vol. 37, pp. 151–182.
Dowling, A. P., and Stow, S. R. (2005), “Acoustic Analysis of Gas Turbine Combustors,”
Combustion Instabilities in Gas Turbine Engines, edited by T. Lieuwen, and V. Yang,
Vol. 210, Progress in Astronautics and Aeronautics, pp. 369–414.
Doyle, J. C., Francis, B. A., and Tannenbaum, A. R. (1992), Feedback Control Theory,
Macmillan, New York.
Drazin, P., and Reid, W. (1981), Hydrodynamic Stability, Cambridge University Press,
Cambridge, UK.
Dryden, H. L. (1955) “Fifty Years of Boundary-Layer Theory and Experiment,” Science,
Vol. 121, pp. 375–380.
Dubief, Y., and Lele, S. K. (2001), “Direct Numerical Simulation of Polymer Flow,” Annual
Research Briefs, Center for Turbulence Research, NASA Ames/Stanford Univ.,
pp. 197–208.
Ducruix, S., and Candel, S. (2004), “External Flow Modulation in Computational Fluid
Dynamics,” AIAA Journal, Vol. 42, pp. 1550–1558.
460 REFERENCES

Ducruix, S., Schuller, T., Durox, D., and Candel, S. (2003), “Combustion Dynamics and
Instabilities: Elementary Coupling and Driving Mechanisms,” Journal of Propulsion
and Power, Vol. 19, No. 3, pp. 722–734.
Ducruix, S., Schuller, T., Durox, D., and Candel, S. (2005), “Combustion Instability
Mechanisms in Premixed Combustors,” Combustion Instabilities in Gas Turbine
Engines, edited by T. Lieuwen, and V. Yang, Vol. 210, Progress in Astronautics and
Aeronautics, pp. 179–212.
Dullerud, G. E., and Paganini, F. (2004), A Course in Robust Control Theory: A Convex
Approach, Texts in Applied Mathematics, Vol. 36, Springer, New York.
Eggenspieler, G., and Menon, S. (2004), “Large-Eddy Simulation of Pollutant Emission
from DOE-HAT Combustor,” Journal of Propulsion and Power, Vol. 20, No. 6,
pp. 1076–1086.
Eggenspieler, G., and Menon, S. (2005), “Combustion and Emission Modeling near Lean
Blow-Out in Gas Turbine Engines,” Progress in Computational Fluid Dynamics, Vol. 5,
No. 6, pp. 281–297.
Ekaterinaris, J. A. (2004), “Prediction of Active Flow Control Performance on Airfoils and
Wings,” Aerospace Science and Technology, Vol. 8, pp. 401–410.
Eldredge, R., and Bons, J. (2004), “Active Control of a Separating Boundary Layer with
Steady Vortex Generating Jets,” AIAA Paper 2004-0751.
Enloe, C. L., McLaughlin, T. E., VanDyken, R. D., Kachner, K. D., Jumper, E. J., and
Corke, T. C. (2003), “Mechanisms and Responses of a Single Dielectric Barrier
Plasma,” AIAA Paper 2003-1021.
Enloe, C. I., McLaughlin, T. E., VanDyken, R. D., Kachner, K. D., Jumper, E. J., Corke,
T. C., Post, M., and Haddad, O. (2004), “Mechanisms and Responses of a Single
Dielectric Barrier Plasma Actuator: Geometric Effects,” AIAA Journal, Vol. 42, No. 3,
pp. 595–604.
Enloe, C. L., McLaughlin, T., Font, G. I., and Baughn, J. W. (2006), “Parameterization of
Temporal Structure in the Single-Dielectric-Barrier Aerodynamic Plasma Actuator,”
AIAA Journal, Vol. 44, No. 6, pp. 1127–1136.
Eschricht, D., Jordan, P., Wei, M., Freund, J., and Thiele, F. (2007), “Analysis of Noise-
Controlled Shear Layers,” AIAA Paper 2007-3660.
Everson, R., and Sirovich, L. (1995), “Karhunen-Loève Procedure for Gappy Data,”
Journal of the Optical Society of America A, Vol. 12, No. 8, pp. 1657–1664.
Fadlun, E. A., Verzicco, R., Orlandi, P., and Mohd-Yusof, J. (2000), “Combined Immersed-
Boundary Finite-Difference Methods for Three-Dimensional Complex Flow Simula-
tions,” Journal of Computational Physics, Vol. 161, pp. 35–60.
Farge, M. (1992), “Wavelet Transforms and their Applications to Turbulence,” Annual
Review of Fluid Mechanics, Vol. 24, pp. 395–458.
Farhood, M., Beck, C. L., and Dullerud, G. E. (2005), “Model Reduction of Periodic
Systems: a Lifting Approach,” Automatica, Vol. 41, pp. 1085–1090.
Farrell, B. F. (1988), “Optimal Excitation of Perturbations in Viscous Shear Flow,” Physics
of Fluids, Vol. 31, pp. 2093–2102.
Farrell, B. F., and Ioannou, P. J. (1996), “Generalized Stability Theory. Part I: Autonomous
Operators,” Journal of the Atmospheric Sciences, Vol. 53, No. 14, pp. 2025–2040.
Fasel, H. F., Seidel, J., and Wernz, S. (2002), “A Methodology for Simulations of Complex
Turbulent Flows,” Journal of Fluids Engineering, Vol. 124, No. 4, pp. 933–942.
Fasel, H. F., Gross, A., and Wernz, S. (2006), “Investigation of Turbulent Coanda Wall Jets
Using DNS and RANS,” Applications of Circulation Control Technology, edited by
REFERENCES 461

R. D. Joslin, and G. S. Jones, Vol. 214, AIAA Progress in Aeronautics and Astronautics
Series, Reston, VA, pp. 401–420.
Federspiel, J. F., Bangert, L. S., Wing, D. J., and Hawkes, T. (1995), “Fluidic Control of
Nozzle Flow—Some Performance Measurements,” AIAA Paper 95-2605.
Fedorov, A. V., Malmuth, N. D., Rasheed, A., and Hornung, H. G., (2001), “Stabilization
of Hypersonic Boundary Layers by Porous Coatings,” AIAA Journal, Vol. 39, No. 4,
pp. 605–610.
Fedorov, A., Shiplyuk, A., Maslov, A., Burov, E., and Malmuth, N. D. (2003), “Stabilization
of a Hypersonic Boundary Layer Using an Ultrasonically Absorptive Coating,” Journal
of Fluid Mechanics, Vol. 479, pp. 99–124.
Fedorov, A. V., Shiplyuk, A., Maslov, A., Kozlov, V., Sidorenko, A., and Malmuth, N. D.
(2006), “Hypersonic Laminar Flow Control Using a Porous Coating of Random
Microstructure,” AIAA Paper 2006-1112.
Fernholtz, H. H., Janke, G., Schober, M., Wagner, P. M., and Warnack, D. (1996), “New
Developments and Applications of Skin-Friction Measuring Techniques,” Measure-
ment Science Technology, Vol. 7, No. 10, pp. 1396–1409.
Ffowcs Williams, J. E., and Hawkings, D. L. (1969), “Sound Generated by Turbulence and
Surfaces in Arbitrary Motion,” Philosophical Transactions of the Royal Society,
Vol. A264, pp. 321–342.
Ffowcs Williams, J. E., and Kempton, A. J. (1978), “The Noise from the Large-Scale
Structure of a Jet,” Journal of Fluid Mechanics, Vol. 84, pp. 673–698.
Fiedler, H. E., Nayeri, C., Spieweg, R., and Paschereit, C. O. (1998), “Three-Dimensional
Mixing Layers and Their Relatives,” Experimental Thermal and Fluid Science, Vol. 16,
pp. 3–21.
Filz, C., Lee, D., Orkwis, P. D., and Turner, M. G. (2003), “Modeling of Two Dimensional
Directed Synthetic Jets Using Neural Network-Based Deterministic Source Terms,”
AIAA Paper 2003-3456.
Fischer, F. A. (1955), Fundamentals of Electroacoustics, Interscience Publishers, New York.
Fisher, D. F., Cobleigh, B. R., Banks, D. W., Hall, R. M., and Wahls, R. A. (1998), “Reynolds
Number Effects at High Angles of Attack,” AIAA Paper 98-2879.
Flamm, J. D. (1998), “Experimental Study of a Nozzle Using Fluidic Counterflow for
Thrust Vectoring,” AIAA Paper 98-3255.
Flamm, J. D. (1996), “Internal Performance Characteristics of a Nozzle Using Fluidic
Counterflow for Thrust Vectoring,” M.S. Thesis, George Washington University,
Washington DC.
Flamm, J. D., Deere, K. A., Berrier, B. L., and Johnson, S. K. (2005), “An Experimental
Study of a Dual Throat Fluidic Thrust Vectoring Nozzle Concept,” AIAA Paper 2005-
3503.
Flamm, J. D., Deere, K. A., Mason, M. L., Berrier, B. L., and Johnson, S. K. (2006),
“Design Enhancements of the Two-Dimensional, Dual Throat Fluidic Thrust Vectoring
Nozzle Concept,” AIAA Paper 2006-3701.
Flamm, J. D., Deere, K. A., Berrier, B. L., and Johnson, S. K. (2007), “Experimental Study
of an Axisymmetric Dual Throat Fluidic Thrust Vectoring Nozzle for a Supersonic
Business Jet,” AIAA Paper 2007-5084.
Foss, J. K., and Zaman, K. B. M. Q. (1999), “Large- and Small-Scale Vortical Motions in a
Shear Layer Perturbed by Tabs,” Journal of Fluid Mechanics, Vol. 382, pp. 307–329.
Fourguette, D., Modarress, D., Taugwalder, F., Wilson, D., Koochesfahani, M., and Gharib,
M. (2001), “Miniature and MOEMS Flow Sensors,” AIAA Paper 2001-2982.
462 REFERENCES

Franke, R. (1982), “Scattered Data Interpolation: Tests of Some Methods,” Mathematics of


Computation, Vol. 38, pp. 181–200.
Franklin, G., Powell, J. D., and Emami-Naeini, A. (2005), Feedback Control of Dynamic
Systems, 5th ed., Prentice-Hall, NJ.
Fransson, J. H. M., Brandt, L., Talamelli, A., and Cossu, C. (2005), “Experimental Study
of Stabilization of Tollmien-Schlicting Waves by Finite Amplitude Streaks,” Physics of
Fluids, Vol. 17, 054110.
Freund, J. B., and Moin, P. (2000), “Jet Mixing Enhancement by High-Amplitude Fluidic
Actuation,” AIAA Journal, Vol. 38, No. 10, pp. 1863–1870.
Freund, J. B., Bodony, D. J., and Lele, S. K. (2002), “Turbulence Interactions Leading to
Far-Field Jet Noise,” Proceedings of the Summer Program, Center for Turbulence
Research, pp. 15–25.
Freymuth, P. (1966), “On Transition in a Separated Laminar Boundary Layer,” Journal of
Fluid Mechanics, Vol. 25, pp. 683–704.
Fuchs, H. V. (1972), “Space Correlations of the Fluctuating Pressure in Subsonic Turbulent
Jets,” Journal of Sound and Vibration, Vol. 23, No. 1, pp. 77–99.
Fuchs, H. V., and Michel, U. (1978), “Experimental Evidence of Turbulent Source
Coherence Affecting Jet Noise,” AIAA Journal, Vol. 16, pp. 871–872.
Gad-el-Hak, M. (1996), “Modern Developments in Flow Control,” Applied Mechanics
Reviews, Vol. 48, pp. 365–379.
Gad-el-Hak, M. (1998), “Introduction to Flow Control,” Flow Control: Fundamentals and
Practices, edited by M. Gad-el-Hak, A. Pollard, and J. Bonnet, Springer, Berlin,
pp. 199–273.
Gad-el-Hak, M. (2000), “Flow Control: Passive, Active, and Reactive Flow Management,”
Cambridge University Press, London.
Gad-el-Hak, M. (ed.) (2001), The CRC Handbook of MEMS, CRC Press, Boca Raton, FL.
Gad-el-Hak, M., Pollard, A., and Bonnet, J. (eds.) (1998), Flow Control: Fundamentals
and Practices, Springer, Berlin.
Gaitonde, D., Shang, J., and Young, J. (1999), “Practical Aspects of Higher-Order Numerical
Schemes for Wave Propagation Phenomena,” International Journal for Numerical
Methods in Engineering, Vol. 45, pp. 1849–1869.
Gaitonde, D. V., Visbal, M. R., and Roy, S. (2005), “Control of Flow Past a Wing Section
with Plasma-Based Body Forces,” AIAA Paper 2005-5302.
Gallaire, F., Chomaz, J.-M., and Huerre, P. (2004) “Closed-Loop Control of Vortex
Breakdown: A Model Study,” Journal of Fluid Mechanics, Vol. 511, pp. 67–93.
Gallas, Q., Mathew, J., Kaysap, A., Holman, R., Nishida, T., Carroll, B., Sheplak, M., and
Cattafesta, L. (2002), “Lumped Element Modeling of Piezoelectric-Driven Synthetic
Jet Actuators,” AIAA Paper 2002-125.
Gallas, Q., Holman, R., Nishida, T., Carroll, B., Sheplak, M., and Cattafesta, L. (2003),
“Lumped Element Modeling of Piezoelectric-Driven Synthetic Jet Actuators,” AIAA
Journal, Vol. 41, No. 2, pp. 240–247.
Gaster, M. (1992), “Stability of Velocity Profiles with Reverse Flow,” Instability, Transition
and Turbulence, edited by M. Y. Hussaini and C. L. Streett, pp. 212–215.
Gaster, M., Kit, E., and Wygnanski. I. (1985), “Large-Scale Structures in a Forced Turbulent
Mixing Layer,” Journal of Fluid Mechanics, Vol. 150, pp. 23–39.
Gatski, T. B., and Rumsey, C. L. (2002), “Linear and Nonlinear Eddy Viscosity Models,”
Closure Strategies for Turbulent and Transitional Flows, edited by B. E. Launder and
N. D. Sandham, Cambridge University Press, Cambridge, UK, pp. 9–46.
REFERENCES 463

Gault, D. E. (1949), “Boundary-Layer and Stalling Characteristics of the NACA 63-009


Airfoil Section,” NACA TN-1894.
Gault, D. E. (1955), “An Experimental Investigation of Regions of Separated Laminar
Flow,” NACA-TN-3505.
Gelb, A. (ed.) (1974), Applied Optimal Estimation. MIT Press, Cambridge, MA.
George, W. K., Wanstrom, M., and Jordan, P. (2007), “Identifying Aeroacoustic Sources,”
AIAA Paper 2007-3603.
Germano, M., Piomelli, U., Moin, P., and Cabot, W. H. (1991), “A Dynamic Subgrid-Scale
Eddy-Viscosity Model,” Physics of Fluids A, Vol. 3, No. 7, pp. 1760–1765.
Geurts, B. J. (1999), “Balancing Errors in LES,” Direct and Large-Eddy Simulation III,
edited by P. R. Voke, N. D. Sandham, and L. Kleiser, Kluwer Academic Publishers,
Dordrecht, pp. 1–12.
Ghias, R., Mittal, R., and Lund, T. S. (2004), “A Non-Body Conformal Grid Method for
Simulation of Compressible Flows with Complex Immersed Boundaries,” AIAA Paper
2004-0080.
Ghoniem, A., Annaswamy, A., Park, S., and Sobhani, Z. (2005), “Stability and Emissions
Control Using Air Injection and H2 Addition in Premixed Combustion,” Proceedings of
the Combustion Institute, Vol. 30, pp. 1765–1773.
Giannakoglou, K. (2004), Optimization Methods and Tools for Multicriteria, Multi-
disciplinary Design, Application to Aeronautics and Turbomachinery, VKI Lecture
Series, Von Karman Inst., Belgium.
Giannetti, F., and Luchini, P. (2007), “Structural Sensitivity of the First Instability of the
Cylinder Wake,” Journal of Fluid Mechanics, Vol. 581, pp. 167–197.
Gibb, J., and Anderson, B. H. (1995), “Vortex Flow Control Applied to Aircraft Intake
Ducts,” Proceedings of the Royal Aeronautical Society, Conf. Paper No. 14.
Gilarranz, J. L., Traub, L. W., and Rediniotis, O. K. (2005), “A New Class of Synthetic Jet
Actuators—Part I: Design, Fabrication and Bench Top Characterization,” Journal of
Fluids Engineering, Vol. 127, No. 2, pp. 367–376.
Giles, M. B., and Pierce, N. A. (2001a), “Analytic Adjoint Solutions for the Quasi-One-
Dimensional Euler Equations,” Journal of Fluid Mechanics, Vol. 426, pp. 327–345.
Giles, M. B., and Pierce, N. A. (2001b), “An Introduction to the Adjoint Approach to
Design,” Flow, Turbulence and Combustion, Vol. 65, pp. 393–415.
Ginevsky, A. S., Vlasov, Y. V., and Karavosov, R. K. (2004), “Acoustic Control of Turbulent
Jets,” Foundations of Engineering Mechanics, Springer-Verlag, Berlin.
Girimaji, S. S., and Lavin, T. A. (2006), “Investigation of Turbulent Square Jet Using
PANS Method,” AIAA Paper 2006-0488.
Giuliano, V. J., and Wing, D. J. (1997), “Static Investigation of a Fixed-Aperture Exhaust
Nozzle Employing Fluidic Injection for Multiaxis Thrust Vector Control,” AIAA Paper
97-3149.
Glauert, M. B. (1947), “The Application of the Exact Method of Airfoil Design,”
Aeronautical Research Council R. & M. No. 2683, London.
Glauert, M. B., Walker, W. S., Raymer, W. G., and Gregory, N. (1948), “Wind-Tunnel Tests
on a Thick Suction Aerofoil with a Single Slot,” Aeronautical Research Council R. &
M. No. 2646.
Glauser, M., and Walker, S. (1998), “Active Flow Control Technology to Cut Millions from
Jet Engine Life Cycle Costs,” AFRL Research Highlights, pp. 2–4.
Glezer, A. (2008), “Dynamic Flight Maneuvering Using Trapped Vorticity Flow Control,”
AIAA Paper 2008-0522.
464 REFERENCES

Glezer, A., and Amitay, M. (2002), “Synthetic Jets,” Annual Review of Fluid Mechanics,
Vol. 34, pp. 503–529.
Glezer, A., Amitay, M., and Honohan, A. (2005), “Aspects of Low- and High-Frequency
Actuation for Aerodynamic Flow Control,” AIAA Journal, Vol. 43, pp. 1501–1511.
Godard, G., and Stanislas, M. (2006), “Control of a Decelerating Boundary Layer. Part 3:
Optimization of Round Jets Vortex Generators,” Aerospace Science and Technology,
Vol. 10, pp. 455–464.
Göksel, B., and Rechenberg, I. (2004), “Active Separation Flow Control Experiments in
Weakly Ionized Air,” Advances in Turbulence X: Proceedings of the 10th Euromech
European Turbulence Conference, edited by H. I. Andersson and P.-Å. Krogstad,
CIMNE, Barcelona.
Göksel, B., Greenblatt, D., Rechenberg I., Nayeri, C. N., and Paschereit, C. O. (2006),
“Steady and Unsteady Plasma Wall Jets for Separation and Circulation Control,” AIAA
Paper 2006-3686.
Göksel, B., Greenblatt, D., Rechenberg, I., Kastantin, Y., Nayeri, C. N., and Paschereit, C.
O. (2007), “Pulsed Plasma Actuators for Active Flow Control at MAV Reynolds
Numbers,” Notes on Numerical Fluid Mechanics and Multidisciplinary Design,
Springer, Berlin, Chap. 16, pp. 42–65.
Goldberg, U., Peroomian, O., and Chakravarthy, S. (1998), “A Wall-Distance-Free k-e
Model with Enhanced Near-Wall Treatment,” Journal of Fluids Engineering, Vol. 120,
pp. 457–462.
Goldschmied, F. R. (1981), “Wind Tunnel Demonstration of an Optimized LTA System
with 65% Power Reduction and Neutral Static Stability,” AIAA Paper 83-1981.
Goldschmied, F. R. (1987), “Fuselage Self-Propulsion by Static Pressure Thrust: Wind
Tunnel Verification,” AIAA Paper 87-2935.
Goldstein, M. E. (1976), Aeroacoustics, McGraw-Hill, New York, p. 253.
Goldstein, M. E. (1981), “The Coupling between Flow Instabilities and Incident
Disturbances at a Leading-Edge,” Journal of Fluid Mechanics, Vol. 104, pp. 217–246.
Goldstein, M. E. (1983), “The Evolution of Tollmien–Schlichting Waves Near a Leading
Edge,” Journal of Fluid Mechanics, Vol. 127, pp. 59–81.
Goldstein, M. E. (2005), “On Identifying the True Sources of Aerodynamic Sound,”
Journal of Fluid Mechanics, Vol. 526, pp. 337–347.
Goldstein, D., Handler, R., and Sirovich, L. (1993), “Modelling a No-Slip Flow
Boundary with an External Force Field,” Journal of Computational Physics, Vol. 105,
pp. 354–366.
Golub, G., and Loan, C. V. (1996), Matrix Computations, 3rd ed., Johns Hopkins Univ.
Press, Baltimore, MA.
Goodson, K. (1966), “Longitudinal Aerodynamic Characteristics of a Flapped Tilt-Wing
Four-Propeller V/STOL Transport Model,” NASA TN-D-3217.
Gorton, S. A., Owens, L. R., Jenkins, L. N., Allan, B. G., and Schuster, E. P. (2004), “Active
Flow Control on a Boundary-Layer-Ingesting Inlet,” AIAA Paper 2004-1203.
Gostelow, J. P., and Thomas, R. L. (2003), “Response of a Laminar Separation Bubble to
an Impinging Wake,” Proceedings of ASME Turbo Expo 2003: Power for Land, Sea,
and Air, GT2003–38972.
Gottlieb, D., and Orszag, S. A. (1977), Numerical Analysis of Spectral Methods: Theory
and Applications, CBMS-NSF Regional Conference Series in Applied Mathematics
26, SIAM, Philadelphia, PA.
REFERENCES 465

Goudsblom, U. (1986), “The Human Monopoly on the use of Fire: its Origins and
Conditions,” Human Volution, Vol. 1, pp. 517–523.
Greatrex, F. B. (1954), “Engine Noise,” Journal of the Royal Aeronautical Society, Vol. 58,
p. 223.
Green, S. I. (1995), Introduction to Vorticity, in Fluid Vortices, edited by S. I. Green, Kluwer
Academic Press, Dordrecht, Chap. 1.
Greenblatt, D. (2006), “Managing Flap Vortices via Separation Control,” AIAA Journal,
Vol. 44, No. 11, pp. 2755–2764.
Greenblatt, D. (2007), “Dual Location Separation Control on a Semi-Span Wing,” AIAA
Journal, Vol. 45, No. 8, pp. 1848–1860.
Greenblatt, D., and Wygnanski, I. (1998), “Dynamic Stall Control by Oscillatory Forcing,”
AIAA Paper 98-0676.
Greenblatt, D., and Wygnanski, I. (2000), “The Control of Flow Separation by Periodic
Excitation,” Progress in Aerospace Sciences, Vol. 36, No. 7, pp. 487–545.
Greenblatt D., and Wygnanski, I. (2001a), “Use of Periodic Excitation to Enhance Airfoil
Performance at Low Reynolds Numbers,” Journal of Aircraft, Vol. 38, No. 1,
pp. 190–192.
Greenblatt, D., and Wygnanski, I. (2001b), “Dynamic Stall Control by Periodic Excitation.
Part 1: NACA 0015 Parametric Study,” Journal of Aircraft, Vol. 38, No. 3, pp. 430–438.
Greenblatt, D., and Wygnanski, I. (2002), “Effect of Leading-Edge Curvature and Slot
Geometry on Dynamic Stall Control,” AIAA Paper 2002-3271.
Greenblatt, D., and Wygnanski, I. (2003), “Effect of Leading-Edge Curvature on Airfoil
Separation Control”, Journal of Aircraft, Vol. 40, No. 3, pp. 473–481.
Greenblatt, D., Darabi, A., Nishri, B., and Wygnanski, I. (1998), “Separation Control by
Periodic Addition of Momentum with Particular Emphasis on Dynamic Stall,”
American Helicopter Society Paper T3-4, Gifu, Japan, April.
Greenblatt, D., Nishri, B., Darabi, A., and Wygnanski, I. (1999), “Some Factors Affecting
Stall Control with Particular Emphasis on Dynamic Stall,” AIAA Paper 99-3504.
Greenblatt, D., Nishri, B., Darabi, A., and Wygnanski, I. (2001a), “Dynamic Stall Control
by Periodic Excitation. Part 2: Mechanisms,” Journal of Aircraft, Vol. 38, No. 3,
pp. 439–447.
Greenblatt, D., Neuburger, D., and Wygnanski, I. (2001b), “Dynamic Stall Control by
Intermittent Periodic Excitation,” Journal of Aircraft, Vol. 38, No. 1, pp. 188–190.
Greenblatt, D., Melton, L., Yao, C., and Harris, J. (2005), “Active Control of a Wing Tip
Vortex,” AIAA Paper 2005-4851.
Greenblatt, D., Paschal, K., Yao, C.-S., and Harris, J. (2005), “A Separation Control CFD
Validation Test Case: Part 2. Zero Efflux Oscillatory Blowing,” AIAA Paper
2005-0485.
Greenblatt, D., Paschal, K. B., Yao, C.-S., Harris, J., Schaeffler, N. W., and Washburn, A. E.
(2004), “A Separation Control CFD Validation Test Case Part 1: Baseline & Steady
Suction,” AIAA Paper 2004-2220.
Greenblatt, D., Paschal, K. B., Yao, C.-S., Harris, J., Schaeffler, N. W., and Washburn, A. E.
(2006a), “Experimental Investigation of Separation Control Part 1: Baseline and Steady
Suction,” AIAA Journal, Vol. 44, No. 12, pp. 2820–2830.
Greenblatt, D., Paschal, K. B., Yao, C.-S., and Harris, J. (2006b), “Experimental Investigation
of Separation Control Part 2: Zero Mass-Efflux Oscillatory Blowing,” AIAA Journal,
Vol. 44, No. 12, pp. 2831–2845.
466 REFERENCES

Greenblatt, D., Göksel, B., Rechenberg, I., Schüle, C., Romann, D., Paschereit, C. O.
(2008), “Dielecttic Barrier Discharge Flow Control at Very Low Flight Reynolds
Numbers,” AIAA Journal, Vol. 46, No. 6, pp. 1528–1541.
Gregory, N., and Walker, W. S. (1952), “Wind-tunnel Tests on the NACA 63A009 Aerofoil
with Distributed Suction over the Nose,” Aeronautical Research Council R. & M. No.
2900.
Gregory, N., Walker, W. S., and Devereux, A. N. (1948), “Wind-Tunnel Tests on the 30
Percent Symmetrical Griffith Airfoil with Distributed Suction over the Nose,”
Aeronautical Research Council R. & M. No. 2647.
Gregory, N., Pankhurst, R. C., and Walker, W. S. (1950), “Wind-tunnel Tests on the
Prevention of Boundary-layer Separation by Distributed Suction at the Rear of a Thick
Airfoil (NPL 153),” Aeronautical Research Council R. & M. No. 2788.
Greska, B. (2005), Supersonic Jet Noise and its Reduction Using Microjet Injection, Ph.D.
Thesis, Florida State Univ., FL.
Greska, B., Krothapalli, A., Seiner, J. M., Jansen, B., and Ukeiley, L. (2005), “The Effect
of Microjet Injection on an F404 Jet Engine,” AIAA Paper 2005-3047.
Gridley, M. C., and Walker, S. H. (1996), “Inlet and Nozzle Tech. for 21st Century Aircraft,”
ASME 96-GT-244.
Griffin, O. M. (1989), “Flow Similitude and Vortex Lock-on in Bluff Body near Wakes,”
Physics of Fluids A, Vol. 1, No. 4, pp. 697–703.
Griffin, L. W., Huber, F. W., and Sharma, O. P. (1996), “Performance Improvement
Through Indexing of Turbine Airfoils: Part II—Numerical Simulation,” Journal of
Turbomachinery, Vol. 118, No. 4, pp. 636–642.
Gross, A., and Fasel, H. F. (2002), “High-Order WENO Schemes Based on the Roe
Approximate Riemann Solver,” AIAA Paper 2002-2735.
Gross, A., and Fasel, H. F. (2005a), “Simulation of Active Flow Control for a Low Pressure
Turbine Blade Cascade,” AIAA Paper 2005-0869.
Gross, A., and Fasel, H. F. (2005b), “Numerical Investigation of Low-Pressure Turbine
Blade Separation Control,” AIAA Journal, Vol. 43, No. 12, pp. 2514–2525.
Gross, A., and Fasel, H. F. (2005c), “Turbulence Modeling for Low Pressure Turbine
Blades,” AIAA Paper 2005–5922.
Gross, A., and Fasel, H. F. (2006a), “Coanda Wall Jet Calculations Using One- and
Two-Equation Turbulence Models,” AIAA Journal, Vol. 44, No. 9, pp. 2095–2107.
Gross, A., and Fasel, H. F. (2006b), “RANS, URANS, and LES of Coanda Wall Jet Flows,”
AIAA Paper 2006-3371.
Gross, A., and Fasel, H. F. (2007a), “Investigation of Low-Pressure Turbine Separation
Control,” AIAA Paper 2007-0520.
Gross, A., and Fasel, H. F. (2007b), “Characteristic Ghost Cell Boundary Condition,” AIAA
Journal, Vol. 45, No. 1, pp. 302–306.
Gross, A., and Fasel, H. F. (2007c), “Self-Adaptive Closed-Loop Control of Low-Reynolds
Number Laminar Separation,” AIAA Paper 2007-3913.
Gross, A., and Fasel, H. F. (2007d), “Control-Oriented Proper Orthogonal Decomposition
Models for Unsteady Flows,” AIAA Journal, Vol. 45, No. 4, pp. 814–827.
Grossman, K., Bohdan, C., and VanWie, D. (2003), “Sparkjet Actuators for Flow Control,”
AIAA Paper 2003-0057.
Grossman, K., Cybyk, B., VanWie, D., and Rigling, M. (2004), “Characterization of
SparkJet Actuators for Flow Control,” AIAA Paper 2004-89.
REFERENCES 467

Guckenheimer, J., and Holmes, P. J. (2002), Nonlinear Oscillations, Dynamical Systems,


and Bifurcations of Vector Fields, 6th ed., Vol. 42, Applied Mathematical Sciences,
Springer, New York.
Guitton, A., Jordan, P., Laurendeau, E., and Delville, J. (2007), “Velocity Dependence of
the Near Pressure Field of Subsonic Jets: Understanding the Associated Source
Mechanisms,” AIAA Paper 2007-3661.
Gunzberger, M. D. (2003), “Perspectives in Flow Control and Optimization,” Advances in
Design and Control, SIAM, Philadelphia, PA.
Guo, D., Cary, A. W., and Agarwal, R. K. (2003), “Numerical Simulation of Vectoring of a
Primary Jet with a Synthetic Jet,” AIAA Journal, Vol. 41, No. 12, pp. 2364–2370.
Gupta, A. K. (1997), “Gas Turbine Combustion: Prospects and Challenges,” Energy
Conversion and Management, Vol. 38, No. 10, pp. 1311–1318.
Gupta, A. K., and Lilley, D. G. (1994), “Combustion and Environmental Challenges for
Gas Turbines in the 1990s,” Journal of Propulsion and Power, Vol. 10, No. 2,
pp. 137–149.
Gursul, I., Srinivas, S., and Batta, G. (1995), “Active Control of Vortex Breakdown over a
Delta Wing,” AIAA Journal, Vol. 33, pp. 1743–1745.
Gushchin, V. R., and Fedorov, A. V. (1989), “Asymptotic Analysis of Inviscid Perturbations
in a Supersonic Boundary Layer,” Zhurnal Prikl. Mekh. i Tekh. Fiz., Vol. 1, pp. 69–75.
Gustafsson, K. M. B., and Johansson, T. G. (2003), “Turbulence Velocity Fields of Slanted
Jets in Cross-flow—Measurements and CFD Simulations,” Turbulence, Heat and Mass
Transfer 4, edited by K. Hanjalic, Y. Nagano, and M. Tummers, Begell House, New
York, pp. 763–770.
Gutmark, E., and Ho, C. (1983), “Preferred Modes and the Spreading Rates of Jets,”
Physics of Fluids, Vol. 26, No. 10, pp. 2932–2938.
Gutmark, E. J., Hanson-Parr, D. M., Parr, T. P., Schadow, K. C., and Wilson, K. J. (1990),
“Use of Chemiluminescence and Neural Networks in Active Combustion Control,”
Proceedings of the Combustion Institute, Vol. 23, pp. 1101–1106.
Gutmark, E. J., Hanson-Parr, D. M., Parr, T. P., and Schadow, K. C. (1991), “Simultaneous
OH and Schlieren Visualization of Premixed Flames at the Lean Blowout Limit,”
Experiments in Fluids, Vol. 12, No. 1, pp. 10–16.
Gutmark, E., Paschereit, C. O., and Weisenstein, W. (1998), “Control of Thermoacoustic
Instabilities and Emissions in an Industrial-Type Gas-Turbine Combustor,” Proceedings
of the Combustion Institute, Vol. 27, pp. 1817–1824.
Guy, Y., Morrow, J., and Mclaughlin, T. (2000), “Parametric Investigation of the Effects
of Active Flow Control on the Normal Force of a Delta Wing,” AIAA Paper
2000-0549.
Guy, Y., Morrow, J., and Mclaughlin, T. (1999), “Control of Vortex Breakdown on a Delta
Wing by Periodic Blowing and Suction,” AIAA Paper 99-132.
Hafenrichter, E., Lee, Y., McIlwain, S., Dutton, J., and Loth, E. (2001), “Experiments on
Normal Shock/Boundary Layer Interaction Control Using Aeroelastic Mesoflaps,”
AIAA Paper 2001-0156.
Halasz, C., Arntz, D., Burdisso, R., and Ng, W. (2005), “Fan Flow Control for Noise
Reduction Part 1: Advanced Trailing Edge Blowing Concepts,” AIAA Paper 2005-3025.
Halfon, E., Nishri, B., Seifert, A., and Wygnanski, I. (2002), “Effects of Elevated Free-
stream Turbulence on Active Control of a Transitional Separation Bubble,” AIAA
Paper 2002-3169.
468 REFERENCES

Halfon, E., Nishri, B., Seifert, A., and Wygnanski, I. (2004), “Effects of Elevated Free-
stream Turbulence on Active Control of a Separation Bubble,” Journal of Fluids
Engineering, Vol. 126, No. 6, pp. 1015–1024.
Hall, J., Pinier, J., Hall, A. M., and Glauser, M. N. (2007), “A Spatio-Temporal
Decomposition of the Acoustic Source in a Mach 0.85 Jet,” AIAA Paper 2007-0442.
Halstead, D. (1996), Boundary Layer Development in Multi-Stage Low Pressure Turbines,
Ph.D. Thesis, Iowa State Univ., Ames, IA.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P., and Shin, H.
(1995), “Boundary Layer Development in Axial Compressors and Turbines, Part 3 of
4: LP Turbines,” ASME Paper 95-GT-463.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P., and Shin, H.-W.
(1997a), “Boundary Layer Development in Axial Compressors and Turbines. Part 1 of
4: Composite Picture,” Journal of Turbomachinery, Vol. 119, No. 1, pp. 114–127.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P., and Shin, H.-W.
(1997b), “Boundary Layer Development in Axial Compressors and Turbines. Part 3 of
4: Turbines,” Journal of Turbomachinery, Vol. 119, No. 2, pp. 225–237.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P., and Shin, H.-W.
(1997c), “Boundary Layer Development in Axial Compressors and Turbines. Part 2 of
4: Compressors,” Journal of Turbomachinery, Vol. 119, No. 3, pp. 426–444.
Halstead, D. E., Wisler, D. C., Okiishi, T. H., Walker, G. J., Hodson, H. P., and Shin H.-W.
(1997d), “Boundary Layer Development in Axial Compressors and Turbines. Part 4 of
4: Computations and Analyses,” Journal of Turbomachinery, Vol. 119, No. 1,
pp. 128–139.
Halstead, D. E. (1997e), “Flowfield Unsteadiness and Turbulence in Multistage Low
Pressure Turbines,” Conference on Boundary Layer Transition in Turbomachines,
Syracuse Univ., Minnowbrook, NY, Sept.
Hammond, E. P., Bewley, T. R., and Moin, P. (1998), “Observed Mechanisms for Turbulence
Attenuation and Enhancement in Opposition-Controlled Wall-Bounded Flows,”
Physics of Fluids, Vol. 10, pp. 2421–2423.
Hamstra, J. W., and Miller, D. N. (2000) “Active Inlet Flow Control Technology
Demonstration,” Aeronautical Journal, Vol. 104, No. 1040.
Hamstra, J. W., Miller, D. N., Truax, P. P., Anderson, B. A., Wendt, B. J. (2000), “Active
Flow Control Technology Demonstration,” Royal Aeronautical Society Aeronautical
Journal, Vol. 104, No. 1040, pp. 473–479.
Han, G., Zhou, M., and Wygnanski, I. (2004), “On Streamwise Vortices and their Role in
the Development of a Curved Wall Jet,” Physics of Fluids, Vol. 18–9, 095104-1-14.
Hanjalic, K., and Jakirlic, S. (2002), “Second-Moment Turbulence Closure Modelling,”
Closure Strategies for Turbulent and Transitional Flows, edited by B. E. Launder and
N. D. Sandham, Cambridge University Press, Cambridge, UK, pp. 47–101.
Hansen, L., and Bons, J. (2006), “Flow Measurements of Vortex Generator Jets in Separating
Boundary Layer,” Journal of Propulsion and Power, Vol. 22, No. 3, pp. 558–566.
Haritonidis, J. H. (1989), “The Measurement of Wall Shear Stress,” Advances in Fluid
Mechanics Measurements, edited by M. Gad-el-Hak, Springer-Verlag, Berlin,
pp. 229–261.
Harrison, S. M., Gutmark, E. J., and Martens, S. (2007), “Jet Noise Reduction by Fluidic
Injection on a Separate Flow Exhaust System,” AIAA Paper 2007-0439.
Harrje, D. T., and Reardon, F. (eds.) (1972), “Liquid Propellant Rocket Combustion
Instability,” NASA SP-194.
REFERENCES 469

Hartmann, J., and Trolle, B. (1927), “A New Acoustic Generator,” Journal of Scientific
Instruments, Vol. 4, No. 4, pp. 101–111.
Hassan, A. A. (2003), “Improving Flap Aerodynamics using Oscillatory Jet Control,”
AIAA Paper 2003-3664.
Hassan, A. A. (2004a), “Oscillatory and Pulsed Jets for Improving Airfoil Aerodynamics –
Numerical Simulation,” AIAA Paper 2004-0227.
Hassan, A. A. (2004b), “Strategies for Tiltrotor Airframe Download Reduction Using
Active Flow Control,” 60th Forum of the American Helicopter Society, June.
Hassan, A. A. (2006), “A Two-Point Active Flow Control Strategy for Improved Airfoil
Stall/Post-Stall Aerodynamics,” AIAA Paper 2006-99.
Hassan, A. A., and Charles, B. D. (1991), “Simulation of Three-Dimensional Blade-Vortex
Interactions Using a Finite-Difference Technique,” Journal of the American Helicopter
Society, Vol. 36, No. 3, pp. 71–83.
Hassan, A. A., and Munts, E. (2000), “Transverse and Near-Tangent Synthetic Jets for
Aerodynamic Flow Control,” AIAA Paper 2000-4334.
Hassan, A. A., Straub, F. K., and Charles, B. D. (1996), “Effects of Surface Blowing/
Suction on the Aerodynamics of Helicopter Rotor Blade–Vortex Interactions (BVI)—
Numerical Simulation,” 52nd Annual Forum of the American Helicopter Society,
Washington, DC, June.
Hassan, A. A., Osborne, B., Schwimley, S., and Billman, G. (2007), “Control of Shock–
Boundary Layer Interactions (SBLIs) Using an Oscillatory Jet,” AIAA Paper
2007-0476.
Hathout, J. P., Fleifil, M., Annaswamy, A. M., and Ghoniem, A. F. (2002), “Combustion
Instability Active Control Using Periodic Fuel Injection,” Journal of Propulsion and
Power, Vol. 18, pp. 390–399.
Haykin, S. S. (1996), Adaptive Filter Theory. 3rd ed., Prentice-Hall.
Haynes, T. S., and Reed, H. L. (2000), “Simulation of Swept-Wing Vortices Using Nonlinear
Parabolized Stability Equations,” Journal of Fluid Mechanics, Vol. 405, pp. 325–349.
He C., and Corke, T. C. (2007), “Numerical and Experimental Analysis of Plasma Flow
Control Over a Hump Model,” AIAA Paper 2007-0935.
He, J. W., Glowinski, R., Metcalfe, R., Nordlander, A., and Periaux, J. (2000), “Active
Control and Drag Optimization for Flow Past a Circular Cylinder,” Journal of
Computational Physics, Vol. 163, pp. 83–117.
Heine, C., Spiess, M. C., Möser, M., and Fiedler, H. E. (1997), “The Influence of Feedback
Control on the Vortex Dynamics in a Turbulent Cylinder Wake,” Euromech Colloquium
361—Active Control of Turbulent Shear Flows, Berlin.
Helmholtz, H. L. F. (1868), “Über discontinuierliche Fluessigkeits-Bewegungen,”
Monatsberichte der Koeniglichen Preussiche Akademie der Wissenshaften zu Berlin,
Vol. 23, p. 215.
Henderson, B., Kinzie, K., Whitmire, J., and Abeysinghe, A. (2005), “The Impact of Fluidic
Chevrons on Jet Noise,” AIAA Paper 2005-2888.
Henderson, B., Kinzie, K., Whitmire, J., and Abeysinghe, A. (2006), “Aeroacoustic
Improvements to Fluidic Chevron Nozzles,” AIAA Paper 2006-2706.
Henning, L., and King, R. (2005), “Drag Reduction by Closed-Loop Control of a Separated
Flow over a Bluff body with a Blunt Trailing Edge,” Proceedings of the 44th IEEE
Conference on Decision and Control, Seville, Spain, pp. 494–499.
Henningson, D. S., and Schmid, P. J. (1992), “Vector Eigenfunction Expansion for Plane
Channel Flows,” Studies in Applied Mathematics, Vol. 87, pp. 15–43.
470 REFERENCES

Herbert, Th. (1984), “Secondary Instability of Shear Flows,” AGARD-R-709 Special


Course on Stability and Transition of Laminar Flow, pp. 7.1–7.13.
Herbert, Th. (1997), “Parabolized Stability Equations,” Annual Review of Fluid Mechanics,
Vol. 29, pp. 245–283.
Hermann, J., Gleis, S., and Vortmeyer, D. (1996), “Active Instability Control of Spray
Combustors by Modulation of the Liquid Fuel Flow Rate,” Combustion Science and
Technology, Vol. 118, Nos. 1–3, pp. 1–25.
Hermann, J., and Hoffmann, S. (2005), “Implementation of Active Control in a Full-Scale
Gas-Turbine Combustor,” Combustion Instabilities in Gas Turbine Engines, edited
by T. Lieuwen, and V. Yang, Vol. 210, Progress in Astronautics and Aeronautics,
pp. 611–634.
Hileman, J. I., Reynolds, T. G., de la Rosa Blanco, E., Law, T. R., and Thomas, S. (2007a),
“Development of Approach Procedures for Silent Aircraft,” AIAA Paper 2007-0451.
Hileman, J. I., Spakovszky, Z. S., Drela, M., and Sargeant, M. A. (2007b), “Airframe
Design for Silent Aircraft,” AIAA Paper 2007-0453.
Hill, D. C. (1992), “A Theoretical Approach for the Restabilization of Wakes,” AIAA Paper
92-0067.
Hill, D. C. (1995), “Adjoint Systems and Their Role in the Receptivity Problem for
Boundary Layers,” Journal of Fluid Mechanics, Vol. 292, pp. 183–204.
Hiller, S. J., and Seitz, P. A. (2006), “The Interaction Between a Fluidic Actuator and Main
Flow Using SAS Turbulence Modeling,” AIAA Paper 2006-3678.
Hirsch, M. W., Smale, S., and Devaney, R. L. (2004), Differential Equations, Dynamical
Systems and an Introduction to Chaos, Academic Press/Elsevier, Amsterdam.
Hites, M., Nagib, H., Bachar, T., and Wygnanski, I. (2001), “Enhanced Performance of
Airfoils at Moderate Mach Numbers Using Zero-Mass Flux Pulsed Blowing,” AIAA
Paper 2001-0734.
Ho, C.-M., and Huang, L.-S. (1982), “Subharmonics and Vortex Merging in Mixing
Layers,” Journal of Fluid Mechanics, Vol. 119, pp. 443–473.
Ho, C. M., and Huerre, P. (1984), “Perturbed Free Shear Layers,” Annual Review of Fluid
Mechanics, Vol. 16, pp. 365–424.
Ho, C.-M., and Tai, Y.-C. (1998), “Microelectromechanical Systems (MEMS) and Fluid
Flows,” Annual Review of Fluid Mechanics, Vol. 30, pp. 579–612.
Hodson, H. P., and Howell, R. J. (2005), “Bladerow Interactions, Transition, and High-Lift
Aerofoils in Low-Pressure Turbines,” Annual Review of Fluid Mechanics, Vol. 37,
pp. 71–98.
Hodson, H., Huntsman, I., and Steele, A. (1993), “An Investigation of Boundary Layer
Development in a Multistage LP Turbine,” ASME Paper 93-GT-310.
Hoepffner, J. (2006), “Stability and Control of Shear Flows Subject to Stochastic
Excitation,” Thesis, KTH Mechanics.
Hoepffner, J., Chevalier, M., Bewley, T. R., and Henningson, D. S. (2005), “State Estimation
in Wall-Bounded Flow Systems. Part 1. Perturbed Laminar Flows,” Journal of Fluid
Mechanics, Vol. 534, pp. 263–294.
Hoepffner, J., Akervik, E., Ehrenstein, U., and Henningson, D. S. (2007), “Optimal Growth,
Model Reduction and Control in a Separated Boundary-Layer Flow Using Global
Eigenmodes,” Journal of Fluid Mechanics, Vol. 579, pp. 305–314.
Hoffmann, S., Judith, H., Krebs, W., and Walz, G. (2002), “Detailed Analysis of the
Acoustic Mode Shapes of an Annular Combustion Chamber,” Journal of Engineering
for Gas Turbines of Power, Vol. 124, pp. 3–9.
REFERENCES 471

Högberg, M., and Henningson, D. (1998), “Secondary Instability of Crossflow Vortices in


Falkner-Skan-Cooke Boundary Layers,” Journal of Fluid Mechanics, Vol. 368,
pp. 339–357.
Högberg, M., and Henningson, D. S. (2002), “Linear Optimal Control Applied to
Instabilities in Spatially Developing Boundary Layers,” Journal of Fluid Mechanics,
Vol. 470, pp. 151–179.
Högberg, M., Bewley, T. R., and Henningson, D. S. (2003), “Linear Feedback Control and
Estimation of Transition in Plane Channel Flow,” Journal of Fluid Mechanics,
Vol. 481, pp. 149–175.
Holden, H., and Babinsky, H. (2003), “Shock/Boundary Layer Interaction Control Using
3D Devices,” AIAA Paper 2003-447.
Holden, H. A., and Babinsky, H. (2007), “Effect of Microvortex Generators on Shock/
Boundary Layer Interactions,” Journal of Aircraft, Vol. 44, No. 1, pp. 170–174.
Holman, R., Utturkar, Y., Mittal, R., Smith, B., and Cattafesta, L. (2005), “A Jet Formation
Criterion for Synthetic Jets,” AIAA Journal, Vol. 43, No. 10, pp. 2110–2116.
Holmes, P., Lumley, J. L., and Berkooz, G. (1996), Turbulence, Coherent Structures,
Dynamical Systems and Symmetry, Cambridge University Press, Cambridge, UK.
Hong, B., Yang, V., and Ray, A. (2000), “Robust Feedback Control of Combustion Instability
with Modeling Uncertainty,” Combustion and Flame, Vol. 120, pp. 91–106.
Hong, B., Ray, A., and Yang, V. (2002), “Wide-Range Robust Control of Combustion
Instability,” Combustion and Flame, Vol. 128, pp. 242–258.
Hotchkiss, J. F., and Martin, J. S. (1997), 500 Years of Golf Balls: History and Collector’s
Guide, Antique Trader Books, Dubuque, IA.
Howe, M. S. (1975), “Contributions to the Theory of Aerodynamic Sound, with Application
to Excess Jet Noise and the Theory of the Flute,” Journal of Fluid Mechanics, Vol. 71,
pp. 625–673.
Howe, M. S. (1998), Acoustics of Fluid Structure Interactions, Cambridge University
Press, Cambridge, UK.
Howe, M. S. (2002), “Noise Generated By A Coanda Wall Jet Circulation Control Device,”
Journal of Sound and Vibration, Vol. 249, No. 4, pp. 679–700.
Howell, R. J., Hodsun, H. P., Schulte, V., Stieger, R. D., Schiffer, H.-P., Haselbach, F., and
Harvey, N. W. (2002), “Boundary Layer Development in the BR710 and BR715 LP
Turbine—the Implementation of High-Lift and Ultra-High-Lift Concepts,” Journal of
Turbomachinery, Vol. 124, No. 3, pp. 385–392.
Hsiao, F.-B., Liu, C.-F., and Shyu, J.-Y. (1990), “Control of Wall Separated Flow by Internal
Acoustic Excitation,” AIAA Journal, Vol. 28, No. 8, pp. 1440–1446.
Huang, L.-S., and Ho, C.-M. (1990), “Small-Scale Transition in a Plane Mixing Layer,”
Journal of Fluid Mechanics, Vol. 210, pp. 475–500.
Huang, L. S., Maestrello, L., and Bryant, T. D. (1987), “Separation Control Over an Airfoil
at High Angles of Attack by Sound Emanating from the Surface,” AIAA Paper 87-1261.
Huang, C. Naguib, A., Soupos, E., and Najafi, K. (2002), “A Silicon Micromachined
Microphone for Fluid Mechanics Research,” Journal of Micromechanics and
Microengineering, Vol. 12, pp. 767–774.
Huang, J., Corke, T. C., and Thomas, F. O. (2002), “Separation Control over Low Pressure
Turbine Blades,” KJ 5, Bulletin of the American Physical Society, Division of Fluid
Dynamics 55th Annual Meeting, Austin, TX, Nov.
Huang, J., Corke, T. C., and Thomas, F. O. (2006a), “Plasma Actuators for Separation
Control of Low-Pressure Turbine Blades,” AIAA Journal, Vol. 44, No. 1, pp. 51–57.
472 REFERENCES

Huang, J., Corke, T. C., and Thomas, F. O. (2006b), “Unsteady Plasma Actuators for
Separation Control of Low-Pressure Turbine Blades,” AIAA Journal, Vol. 44, No. 7,
pp. 1477–1487.
Huber, F. W., Johnson, P. D., Sharma, O. P., Staubach, J. B., and Gaddis, S. W. (1996),
“Performance Improvement Through Indexing of Turbine Airfoils: Part I—Experimental
Investigation,” ASME Journal of Turbomachinery, Vol. 118, No. 4, pp. 630–635.
Huerre, P., and Monkewitz, P. (1985), “Absolute and Convective Instabilities in Free Shear
Layers,” Journal of Fluid Mechanics, Vol. 159, pp. 151–168.
Huerre, P., and Monkewitz, P. A. (1990), “Local and Global Instabilities in Spatially
Developing Flows,” Annual Review of Fluid Mechanics, Vol. 22, pp. 473–537.
Hultgren, L. S., and Ashpis, D. E. (2003), “Demonstration of Separation Delay with Glow-
Discharge Plasma Actuators,” AIAA Paper 2003-1025.
Hunt, J. C. R., Wray, A. A., and Moin, P. (1988), “Eddies, Stream, and Convergence Zones
in Turbulent Flows,” Report CTR-S88, Center For Turbulence Research, Stanford, CA.
Hunter, C. A., and Deere, K. A. (1999), “Computational Investigation of Fluidic Counterflow
Thrust Vectoring,” AIAA Paper 99-2669.
Hunter, P. A., and Johnson, H. I. (1954), “A Flight Investigation of the Practical Problems
Associated with Porous-Leading-Edge Suction,” NACA TN-3062.
Hussain, A. K. M. F. (1983), “Coherent Structures—Reality and Myth,” Physics of Fluids,
Vol. 26, No. 10, pp. 2816–2850.
Hussain, A. K. M. F. (1986), “Coherent Structures and Turbulence,” Journal of Fluid
Mechanics, Vol. 173, pp. 303–356.
Hussain, F., and Hyder, H. S. (1989), “Elliptic Jets. Part 1. Characteristics of Unexcited and
Excited Jets,” Journal of Fluid Mechanics, Vol. 208, pp. 257–320.
Hussain, A. K. M. F., and Reynolds, W. C. (1970), “The Mechanics of an Organized Wave
in Turbulent Shear Flow,” Journal of Fluid Mechanics, Vol. 41, Part 2, pp. 241–258.
Hussain, A. K. M. F., and Reynolds, W. C. (1972), “The Mechanics of an Organized Wave
in Turbulent Shear Flow. Part 2. Experimental Results,” Journal of Fluid Mechanics,
Vol. 54, pp. 241–261.
Iaccarino, G., and Verzicco, R. (2003), “Immersed Boundary Technique for Turbulent Flow
Simulations,” Applied Mechanics Review, Vol. 56, No. 3, pp. 331–347.
Iaccarino, G., Marongiu, C., Catalano, P., and Amato, M. (2004), “RANS Modeling and
Simulations of Synthetic Jets,” AIAA Paper 2004-2223.
Ibrahim, M. K., Kunimura, R., and Nakamura, Y. (2002), “Mixing Enhancement of
Compressible Jets by Using Unsteady Microjets as Actuators,” AIAA Journal, Vol. 40,
No. 4, pp. 681–688.
Ilak M., and Rowley, C. W. (2006), “Reduced-Order Modeling of Channel Flow Using
Traveling POD and Balanced POD,” AIAA Paper 2006-3194.
Ilak M., and Rowley, C. W. (2008), “Modeling of Transitional Channel Flow using Balanced
Proper Orthogonal Decomposition,” Physics of Fluids, Vol. 20, 034103.
Ingard, U., and Ising, H. (1967), “Acoustic Nonlinearity of an Orifice,” Journal of the
Acoustical Society of America, Vol. 42, No. 1, pp. 6–17.
Ingard, U., and Labate, S. (1950), “Acoustic Circulation Effects and the Nonlinear
Impedance of Orifices,” Journal of the Acoustical Society of America, Vol. 22, No. 2,
pp. 211–218.
Israel, D. M., Fasse, E. D., and Fasel, H. F. (2002), “Numerical Simulation of Closed Loop
Active Flow Control of Separation,” AIAA Paper 2002-3282.
REFERENCES 473

Israel, D. M., Postl, D., and Fasel, H. F. (2004), “A Flow Simulation Methodology for
Analysis of Coherent Structures and Flow Control,” AIAA Paper 2004-2225.
Ito, K., and Ravindran, S. S. (1998), “A Reduced-Order Method for Simulation and Control
of Fluid Flows,” Journal of Computational Physics, Vol. 143, pp. 403–425.
Jacob, J., Rivir, R., Carter, C., and Estevadeordal, J. (2004), “Boundary Layer Flow Control
Using AC Plasma Actuators,” AIAA Paper 2004-2128.
Jacobson, S. A., and Reynolds, W. C. (1998), “Active Control of Streamwise Vortices and
Streaks in Boundary Layers,” Journal of Fluid Mechanics, Vol. 360, pp. 179–211.
Jacot, D., and Mabe, J. (2000), “Boeing Active Flow Control System for the V-22,” AIAA
Paper 2000-2473.
Jaimoukha, I. M., and Kasenally, E. M. (1994), “Krylov Subspace Methods for Solving
Large Lyapunov Equations,” SIAM Journal of Numerical Analysis, Vol. 31,
pp. 227–251.
James, R., Jacobs, J., and Glezer, A. (1994), “Experimental Investigation of a Turbulent
Jet Produced by an Oscillating Surface Actuator,” Applied Mechanics Review, Vol. 47,
pp. S127-S131.
Jameson, A. (1988), “Aerodynamic Design via Control Theory,” Journal of Scientific
Computing, Vol. 3, pp. 233–260.
Jameson, A. (1989), “Computational Aerodynamics for Aircraft Design,” Science, Vol. 245,
pp. 361–371.
Jeon, W.-P., and Blackwelder, R. F. (2000), “Perturbations in the Wall Region Using
Flush Mounted Piezoceramic Actuators,” Experiments in Fluids, Vol. 28, No. 6,
pp. 485–496.
Jeong, J., and Hussain, F. (1995), “On the Identification of a Vortex,” Journal of Fluid
Mechanics, Vol. 285, pp. 69–94.
Jimenez, J., and Pinelli, A. (1999), “The Autonomous Cycle of Near-Wall Turbulence,”
Journal of Fluid Mechanics, Vol. 389, pp. 335–359.
Johnson, W. (1988), A Comprehensive Analytical Model of Rotorcraft Aerodynamics and
Dynamics Vol. I: Theory Manual, Vol. II: Users Manual, Johnson Aeronautics, Palo
Alto, CA.
Johnson, C. E., Neumeier, Y., and Zinn, B. T. (2001a), “Demonstration of Active Control of
Combustion Instabilities on a Full-Scale Gas Turbine Combustor,” Proceedings of
ASME Turbo Expo 2001, ASME 2001-GT-0519.
Johnson, C. E., Lieuwen, T., Torres, H. C., and Zinn, B. T. (2001b), “A Mechanism of
Combustion Instability in Lean Premixed Gas Turbine Combustors,” Journal of
Engineering for Gas Turbines and Power, Vol. 123, pp. 182–189.
Johnston, J. P. (1999), “Pitched and Skewed Vortex Generator Jets for Control of Turbulent
Boundary Layer Separation: A Review,” FEDSM99-6917, pp. 1–10.
Johnston, J., and Nishi, M. (1990), “Vortex Generator Jets—a Means for Flow Separation
Control,” AIAA Journal, Vol. 28, No. 6, pp. 989–994.
Jones, I. S. F. (1970), “The Effect of Vortex Generators on the Noise-Producing Region of
a Jet,” Journal of Sound and Vibration, Vol. 11, No. 1, pp. 65–81.
Jones, G. S., and Joslin, R. D., (eds.) (2005), Proceedings of the 2004 NASA/ONR
Circulation Control Workshop, NASA/CP-2005-213509 Parts 1 and 2.
Jones, W. P., and Wille, M. (1996), “Large Eddy Simulation of a Round Jet in a Cross-
Flow,” Engineering Turbulence Modelling and Experiments 3, edited by W. Rodi and
G. Bergeles, Elsevier Sciences, Amsterdam, pp. 199–208.
474 REFERENCES

Jones, C. M., Lee, J. G., and Santavicca, D. A. (1999), “Closed-Loop Active Control of
Combustion Instabilities Using Subharmonic Secondary Fuel Injection,” Journal of
Propulsion and Power, Vol. 15, No. 4, pp. 584–590.
Jordan, P., and Gervais, Y. (2008), “Subsonic Jet Aeroacoustic: Associating Experiment,
Modelling and Simulation,” Experiments in Fluids, Vol. 44, No. 1, pp. 1–21.
Jordan, P., Schlegel, M., Stalnov, O., Noack, B. R., and Tinney, C. E. (2007), “Identifying
Noisy and Quiet Modes in a Jet,” AIAA Paper 2007-3602.
Joseph, D. D. (1976), Stability of Fluid Motions, Vols. I and II, Springer, Berlin.
Joshi, S. S., Speyer, J. L., and Kim, J. (1997), “A Systems Theory Approach to the Feedback
Stabilization of Infinitesimal and Finite-Amplitude Disturbances in Plane Poiseuille
Flow,” Journal of Fluid Mechanics, Vol. 332, pp. 157–184.
Joslin, R. D. (2001), “The Use of DNS in Flow Control,” AIAA Paper 2001-2544.
Joslin, R. D. (2002), “Coupling DNS with Polymer Models for Flow Control,” DNS/
LES—Progress and Challenges, edited by C. Liu, L. Sakell, and T. Beutner,
pp. 157–170.
Joslin, R. D., Gunzburger, M. D., Nicolaides, R. A., Erlebacher, G., and Hussaini, M. Y.
(1995a), “A Self-Contained Automated Methodology for Optimal Flow Control
Validated for Transition Delay,” NASA CR-198215.
Joslin, R. D., Nicolaides, R. A., Erlebacher, G., Hussaini, M. Y., and Gunzburger, M. D.
(1995b), “Active Control of Instabilities in Laminar Boundary-Layer Flow. Part II: Use
of Sensors and Spectral Controller,” AIAA Journal, Vol. 33, No. 8, pp. 1521–1523.
Joslin, R. D., Gunzburger, M. D., Nicolaides, R. A., Erlebacher, G., and Hussaini, M. Y.
(1997),”Self-Contained Automated Methodology for Optimal Flow Control,” AIAA
Journal, Vol. 35, No. 5, pp. 816–824.
Joslin R. D., Thomas R. H., and Choudhari M. M. (2005), “Synergism of Flow and Noise
Control Technologies,” Progress in Aerospace Sciences, Vol. 41, pp. 363–417.
Juve, D., Sunyach, M., and Comte-Bellot, G. (1979), “Filtered Azimuthal Correlations in
the Acoustic Farfield of a Subsonic Jet,” AIAA Journal, Vol. 17, No. 1, pp. 112–113.
Kailasanath, K. (2000), “Review of Propulsion Applications of Detonation Waves,” AIAA
Journal, Vol. 38, No. 9, pp. 1698–1708.
Kailath, T. (1973) “Some New Algorithms for Recursive Estimation in Constant Linear
Systems,” IEEE Transactions of Information Theory, Vol. 19, No. 6, pp. 750–760.
Kalitzin, G., Wu, X., and Durbin, P. A. (2003), “DNS of Fully Turbulent Flow in a
LPT Passage,” International Journal of Heat and Fluid Flow, Vol. 24, No. 4, pp.
636–644.
Kalman, R. E., and Bucy, R. S. (1961), “New Results in Linear Filtering and Prediction
Theory,” ASME Transactions Ser. D: Journal of Basic Engineering, Vol. 83,
pp. 95–107.
Kälvesten, E., Löfdahl, L., and Stemme, G. (1994), “Small Silicon Based Pressure
Transducers for Measurements in Turbulent Boundary Layers,” Experiments in Fluids,
Vol. 17, pp. 24–31.
Kanda, N., Kogoma, M., Jinno, H., Uchiyama, H., and Okazaki, S. (1991), “Atmospheric
Pressure Glows Plasma and its Application to Surface Treatment and Film Deposition,”
Proceedings of the 10th Symposium on Plasma Chemistry, Vol. 3, Paper 3.2–20.
Karagozian, A. R., Cortelezzi, L., and Soldati, A. (2003), “Manipulation and Control of
Transverse Jets,” CISM Courses and Lectures No. 439, Springer-Wein, New York.
Karniadakis, G. E., and Sherwin, S. J. (2005), Spectral/hp Element Methods for CFD,
2nd ed., Oxford University Press, Oxford, UK.
REFERENCES 475

Kasagi, N. (2006), “Toward Smart Control of Turbulent Jet Mixing and Combustion,”
JSME International Journal: Series B Fluids and Thermal Engineering, Vol. 49, No. 4,
pp. 941–950.
Kastner, J., and Samimy, M. (2002), “Development and Characterization of Hartmann
Tube Fluidic Actuators for High-Speed Flow Control,” AIAA Journal, Vol. 40, No. 10,
pp. 1926–1934.
Katz, Y., Nishri, B., and Wygnanski, I. (1989),”The Delay of Turbulent Boundary Layer
Separation by Oscillatory Active Control,” Physics of Fluids, Vol. 1, pp. 179–181.
Kaufmann, A., Nicoud, F., and Poinsot, T. (2002), “Flow Forcing Techniques for
Numerical Simulation of Combustion Instabilities,” Combustion and Flame, Vol. 131,
pp. 371–385.
Kellogg, S. M. (2000), Immersed Boundary Methods with Applications to Flow Control,
M.S. Thesis, Mechanical Engineering and Materials Science, Rice Univ., Houston, TX.
Kelly, M. W. (1956), “Analysis of Some Parameters used in Correlating Blowing-Type
Boundary-Layer Control Data,” NACA R.M. A56F12.
Kelly, R. E. (1967), “On the Resonant Interaction of Neutral Disturbances in Inviscid Shear
Flows,” Journal of Fluid Mechanics, Vol. 31, pp. 789–799.
Kelvin, Lord (William Thomson) (1871), “Hydrokinetic Solutions and Observations,”
Philosophical Magazine, Vol. 42, pp. 362–377.
Kerschen, E. J., and Reba, R. A. (1995), “Active Control of Wake-Airfoil Interaction Noise
by Leading-Edge Actuators,” CEAS/AIAA Paper 95-058.
Kerschen, E. J., Cain, A. B., and Raman, G. (2004), “Analytical Modeling of Helmholtz
Resonator Based Powered Resonance Tubes,” AIAA Paper 2004-2691.
Khalil, H. K. (1996), Nonlinear Systems, 2nd ed., Prentice-Hall, Upper Saddle River, NJ.
Khan, Z. U., and Johnston, J. P. (2000), “On Vortex Generating Jets,” International Journal
of Heat and Fluid Flow, Vol. 21, No. 5, pp. 506–511.
Kibens, V., Dorris, J., Smith, D., and Mossman, F. (1999), “Active Flow Control Technology
Transition: the Boeing ACE Program,” AIAA Paper 99-3507.
Kiedaisch, J., Nagib, H., and Demanett, B. (2006), “Active Flow Control Applied to High-
Lift Airfoils Utilizing Simple Flaps,” AIAA Paper 2006-2856.
Kiedaisch, J., Demanett, B., Reinhard, P., and Nagib, H. (2007), “Active Flow Control for
High Lift Airfoils: Dynamic Flap Actuation,” AIAA Paper 2007-1120.
Kim, J. (2003), “Control of Turbulent Boundary Layers,” Physics of Fluids, Vol. 15,
pp. 1093–1105.
Kim, S.-W., and Benson, T. J. (1992), “Calculation of a Circular Jet in Crossflow with a
Multiple-Time-Scale Turbulence Model,” International Journal of Heat and Mass
Transfer, Vol. 35, No. 10, pp. 2357–2365.
Kim, J., and Bewley, T. R. (2007), “A Linear Systems Approach to Flow Control,” Annual
Review of Fluid Mechanics, Vol. 39, pp. 383–417.
Kim, H. T., Kline, S. J., and Reynolds, W. C. (1971), “The Production of Turbulence near
a Smooth Wall in a Turbulent Boundary Layer,” Journal of Fluid Mechanics, Vol. 50,
pp. 133–160.
Kim, J., Moin, P., and Moser, R. D. (1987), “Turbulence Statistics in Fully Developed
Channel Flow at Low Reynolds Number,” Journal of Fluid Mechanics, Vol. 177,
pp. 133–166.
Kim, K., Lee, J. G., and Santavicca, D. A. (2000), “Effect of Injection Location on the
Effectiveness of an Active Control System Using Secondary Fuel Injection,”
Proceedings of the Combustion Institute, Vol. 28, pp. 739–746.
476 REFERENCES

Kim, B.-H., Williams, D. R., Emo, S., and Acharya, M. (2005), “Modeling Pulsed-Blowing
Systems for Flow Control,” AIAA Journal, Vol. 43, No. 2, pp. 314–325.
King, R., Becker, R., Garwon, M., and Henning, L. (2004), “Robust and Adaptive Closed-
loop Control of Separated Shear Flows,” AIAA Paper 2004-2519.
Kit, E., Wygnanski, I., Friedman, D., Krivonosova, O., and Zhilenko, D. (2007), “On the
Periodically Excited, Plane Turbulent Mixing Layer, Emanating from a Jagged
Partition,” Journal of Fluid Mechanics, Vol. 589, pp. 479–507.
Kjellgren, P., Anderberg, M., and Wygnanski, I. (2000), “Download Alleviation by Periodic
Excitation on a Typical Tilt-Rotor Configuration—Computation and Experiment,”
AIAA Paper 2000-2697.
Kjellgren, P., Cerchie, D., Cullen, L., and Wygnanski, I. (2002), “Active Flow Control on
Bluff Bodies with Distinct Separation Locations,” AIAA Paper 2002-3069.
Kline, S. J., Reynolds, W. C., Schraub, F. A., and Runstadler, P. W. (1967), “The Structure
of Turbulent Boundary Layers,” Journal of Fluid Mechanics, Vol. 30, No. 4,
pp. 741–773.
Knight, M., and Bamber, M. J. (1929), “Wind Tunnel Tests on Aerofoil Boundary Layer
Control using a Backward Opening Slot,” NACA TN-323.
Knight, D., Yan, H., and Zheltovodov, A. (2001), “Large Eddy Simulation of Supersonic
Turbulent Flow in Expansion-Compression Corner,” DNS/LES—Progress and
Challenges, edited by C. Liu, L. Sakell, and T. Beutner, Greyden Press, Columbus, OH,
pp. 183–194.
Knoop, P., Culick, F. E. C., and Zukoski, E. E. (1977), “Extension of the Stability of
Motions in a Combustion Chamber by Nonlinear Active Control Based on Hysteresis,”
Combustion Science and Technology, Vol. 123, pp. 363–376.
Koff, B. L. (1994), “Aircraft Gas Turbine Emissions Challenge,” ASME Journal of
Engineering for Gas Turbine and Power, Vol. 116, No. 3, pp. 474–477.
Konrad, J. (1976), An Experimental Investigation of Mixing in Two-Dimensional Turbulent
Shear Flows with Applications to Diffusion-Limited Chemical Reactions, Ph.D. Thesis,
California Institute of Technology, Pasadena, CA.
Kota, V., and Wright, M. C. M. (2006), “Wake Generator Control of Inlet Flow to Flow
Distortion Noise,” Journal of Sound and Vibration, Vol. 295, No. 1–2, pp. 94–113.
Kotapati, R. B., and Mittal, R. (2005), “Time-Accurate Three-Dimensional Simulations of
Synthetic Jets in Quiescent Air,” AIAA Paper 2005-0103.
Kotapati, R. B., Mittal, R., and Cattafesta, L. N., III (2006), “Numerical Experiments in
Synthetic Jet Based Separation Control,” AIAA Paper 2006-0320.
Kotapati, R. B., Mittal, R., and Cattafesta, L. N., III (2007), “Numerical Study of a
Transitional Synthetic Jet in Quiescent External Flow,” Journal of Fluid Mechanics,
Vol. 581, pp. 287–321.
Kousen, K. A., and Verdon, J. M. (1994), “Active Control of Wake/Blade-Row Interaction
Noise,” AIAA Journal, Vol. 32, No. 10, pp. 1953–1960.
Kovasznay, L. S. G., Kibens, V., and Blackwelder, R. F. (1970), “Large-Scale Motion in the
Intermittent Region of a Turbulent Boundary Layer,” Journal of Fluid Mechanics,
Vol. 41, No. 2, pp. 283–325.
Krishnamoorthy, S., Price, S. J., and Paidoussis, M. P. (2001), “Cross Flow Past an
Oscillating Circular Cylinder: Synchronization Phenomena in the Near Wake,” Journal
of Fluids and Structures, Vol. 15, pp. 955–980.
Krishnan, V., Squires, K. D., and Forsythe, J. R. (2006), “Prediction of Separated Flow
Characteristics over a Hump,” AIAA Journal, Vol. 44, No. 2, pp. 252–262.
REFERENCES 477

Krothapalli, A., Venkatakrishnan, L., Lourenco, L., Greska, B., and Elavarasan, R. (2003),
“Turbulence and Noise Suppression of a High-Speed Jet by Water Injection,” Journal
of Fluid Mechanics, Vol. 491, pp. 131–159.
Krstic, M., and Wang, H.-H. (2000), “Stability of Extremum Seeking Feedback for General
Nonlinear Dynamic Systems,” Automatica, Vol. 36, pp. 595–601.
Kudar, K. L., and Carpenter, P. W. (2007), “Numerical Investigation and Feasibility Study
of a PZT-driven Micro-valve Pulsed-jet Actuator,” Flow, Turbulence and Combustion,
Vol. 78, No. 3–4, pp. 205–222.
Kumar, V., and Alvi, F. S. (2006), “Use of High-Speed Microjets for Active Separation
Control in Diffusers,” AIAA Journal, Vol. 244, No. 2, pp. 273–281.
Kwakernaak, H., and Sivan, R. (1972), Linear Optimal Control Systems, John Wiley,
New York.
Labergue, A., Leger, L., Moreau, E., and Touchard, G. (2005), “Effect of a Plasma Actuator
on an Airflow along an Inclined Wall—P.I.V. and Wall Pressure Measurements,”
Journal of Electrostatics, Vol. 63, No. 6–10, pp. 961–967.
Lachmann, G. V. (ed.) (1961), Boundary Layer and Flow Control, Vols. 1 and 2, Pergamon
Press, New York.
Laible, A., Valsecchi, P., and Fasel, H. (2006), “Numerical Investigation of Secondary
Centrifugal Instabilities in the Coanda Wall Jet,” AIAA Paper 2006-0908.
Lake, J. (1999), Flow Separation Prevention on a Turbine Blade in Cascade at Low
Reynolds Number, Ph.D. Thesis, Air Force Institute of Technology.
Lake, J. P., King, P. I., and Rivir, R. B. (1999), “Reduction of Separation Losses on a
Turbine Blade with Low Reynolds Number,” AIAA Paper 99-0242.
Lal, M., Huggins, J. D., Oljaca, M., Lubarsky, E., Shcherbik, D., Bibik, A., Menon, S., and
Zinn, B. T. (2003a), “Active Control of Instabilities and Emission in High-Pressure
Combustor using Nanomiser Fuel Injector,” AIAA Paper 2003-4936.
Lal, M., Oljaca, M., Lubarsky, E., Shcherbik, D., and Menon, S. (2003b), “Controlling
Combustion Dynamics in a Swirl Combustor via Spray Optimization,” AIAA Paper
2003-4517.
Lal, M., Oljaca, M., Lubarsky, E., Shcherbik, D., Bibik, A., and Menon, S. (2004),
“Controllable Injection for Supercritical Combustion,” AIAA Paper 2004-3383.
Lall, S., Marsden, J. E., and Glavaški, S. (2002), “A Subspace Approach to Balanced
Truncation for Model Reduction of Nonlinear Control Systems,” International Journal
of Robust Nonlinear Control, Vol. 12, pp. 519–535.
Lamb, H. (1932), Hydrodynamics, Dover Publications, New York.
Lang, W., Poinsot, T., and Candel, S. (1987), “Active Control of Combustion Instability,”
Combustion and Flame, Vol. 70, pp. 281–289.
Langford, M. D., Minton, C. M., Ng, W. F., Estevadeordal, J., and Burdisso, R. A. (2005),
“Fan Flow Control for Noise Reduction Part 2: Investigation of Wake-Filling
Techniques,” AIAA Paper 2005-3026.
Langhorne, P. J., Dowling, A. P., and Hooper, N. (1990), “Practical Active Control System for
Combustion Oscillations,” Journal of Propulsion and Power, Vol. 6, No. 3, pp. 324–333.
Larson, E. W., Ratekin, G. H., and O’Connor, G. M. (1981), “Structural Response of the
SSME Fuel Feedline to Unsteady Shock Oscillations,” The Journal of the Acoustical
Society of America, Vol. 70, No. S1, p. S71.
Lasheras, J. C., and Choi, H. (1988), “Three-Dimensional Instability of a Plane Free Shear
Layer: an Experimental Study of the Formation and Evolution of Streamwise Vortices,”
Journal of Fluid Mechanics, Vol. 189, pp. 53–86.
478 REFERENCES

Lau, J. C., Fisher, M. J., and Fuchs, H. V. (1972), “The Intrinsic Structure of Turbulent
Jets,” Journal of Sound and Vibration, Vol. 22, No. 4, pp. 379–406.
Lauga, E., and Bewley, T. R. (2003), “The Decay of Stabilizability with Reynolds Number
in a Linear Model of Spatially Developing Flows,” Proceedings of the Royal Society of
London A, Vol. 459, pp. 2077–2095.
Launder, B., Reece, G., and Rodi, W. (1975), “Progress in the Development of a Reynolds-
Stress Turbulence Closure,” Journal of Fluid Mechanics, Vol. 68, No. 3, pp. 537–566.
Laurendeau, E. (2007), Etude Aeroacoustique d’une Methode de Controle d’un Jet
Subsonique par Fluidevrons, Ph.D. Thesis, Univ. Poitiers, Poitiers, France.
Laurendeau, E., Bonnet, J.-P., Jordan, P., and Delville, J. (2006), “Impact of Fluidic
Chevrons on the Turbulence Structure of a Subsonic Jet,” AIAA Paper 2006-3510.
Leboeuf, R. L., and Mehta, R. D. (1996), “Vortical Structure Morphology in the Initial
Region of a Forced Mixing Layer: Roll-up and Pairing,” Journal of Fluid Mechanics,
Vol. 315, pp. 175–221.
LeConte J. (1858), “On the Influence of Musical Sounds on the Flame of a Jet of Coal
Gas,” The London, Edinburgh and Dublin Philosophical Magazine and Journal of
Science, Fourth Series XV, pp. 235–239.
Lee, C. Y., and Goldstein, D. B. (2001), “DNS of Microjets for Turbulent Boundary Layer
Control,” AIAA Paper 2001-1013.
Lee, C. Y., and Goldstein, D. B. (2002), “Two-Dimensional Synthetic Jet Simulation,”
AIAA Journal, Vol. 40, No. 3, pp. 510–516.
Lee, K. H., Cortelezzi, L., Kim, J., and Speyer, J. (2001), “Application of Reduced-Order
Controller to Turbulent Flows for Drag Reduction,” Physics of Fluids, Vol. 13, No. 5,
pp. 1321–1330.
Lefebvre, A. H. (1995), “The Role of Fuel Preparation in Low-Emission Combustion,”
Journal of Engineering for Gas Turbines and Power, Vol. 117, No. 4, pp. 617–654.
Lefebvre, A. H. (1999), Gas Turbine Combustion, 2nd ed., Taylor and Francis,
Philadelphia, PA.
Leishman, J. G. (2000), Principles of Helicopter Aerodynamics, Cambridge University
Press, Cambridge, UK.
Leitch, T. A., Saunders, C. A., and Ng, W. F. (2000), “Reduction of Unsteady Stator–Rotor
Interaction using Trailing-Edge Blowing,” Journal of Sound and Vibration, Vol. 235,
No. 2, pp. 235–245.
Lele, S. A. (1992), “Compact Finite Difference Schemes with Spectral-Like Resolution,”
Journal of Computational Physics, Vol. 103, No. 1, pp. 16–42.
Lepicovsky, J., Ahuja, K. K., Brown, W. H., and Burrin, R. H. (1985), “Coherent Large-
Scale Structures in High Reynolds Number Supersonic Jets,” NASA CR-3952.
Lesbros, S., Ozawa, T., and Hong, G. (2006), “Numerical Modeling of Synthetic Jets with
Cross Flow in a Boundary Layer at an Adverse Pressure Gradient,” AIAA Paper
2006-3690.
Levenberg, K. (1944), “Solution of Certain Non-Linear Problems in Least Squares,”
Quarterly Applied Mathematics, Vol. 2, pp. 164–168.
Li, F., and Malik, M. R. (1997), “Spectral Analysis of Parabolized Stability Equations,”
Computers and Fluids, Vol. 26, No. 3, pp. 279–297.
Liebeck, R. H. (1976), “On the Design of Subsonic Airfoils for High Lift,” AIAA Paper
76-406.
Liebeck, R. H., and Ormsbee, A. I. (1970) “Optimization of Airfoils for Maximum Lift,”
Journal of Aircraft, Vol. 7, No. 5, pp. 409–416.
REFERENCES 479

Liepmann, H. W., and Nosenchuck, D. M. (1982), “Active Control of Laminar-Turbulent


Transition,” Journal of Fluid Mechanics, Vol. 118, pp. 201–204.
Liepmann, H. W., Brown, G. L., and Nosenchuck, D. M. (1982), “Control of Laminar-
Instability Waves Using a New Technique,” Journal of Fluid Mechanics, Vol. 118,
pp. 187–200.
Lieuwen, T. (2002), “Experimental Investigation of Limit Cycle Oscillations in
an Unstable Gas Turbine Combustor,” Journal of Propulsion and Power, Vol. 18,
pp. 61–67.
Lieuwen, T., and McManus, K. (2002), “That Elusive Hum,” Mechanical Engineering
Power, June, pp. 53–55.
Lieuwen, T., and Yang, V. (eds.) (2005), Combustion Instabilities in Gas Turbine Engines,
Vol. 210, AIAA Progress in Astronautics and Aeronautics, Reston, VA.
Lieuwen, T., and Zinn, B. T. (1998), “The Role of Equivalence Ratio Oscillations in Driving
Combustion Instability in Low NOx Gas Turbines,” Proceedings of the Combustion
Institute, Vol. 27, pp. 1809–1816.
Lieuwen, T., Torres, H., Johnson, C., and Zinn, B. T. (2001), “A Mechanism for Combustion
Instabilities in Premixed Gas Turbine Combustors,” Journal of Engineering for Gas
Turbine and Power, Vol. 123, No. 1, pp. 182–190.
Lighthill, M. J. (1952), “On the Sound Generated Aerodynamically, Part 1: General
Theory,” Proceedings of the Royal Society of London, A211, pp. 554–587.
Lilley, G. F. (1971), “Fourth Monthly Progress Report on Contract F-33615–71-C-1663.
Appendix: Generation of Sound in a Mixing Region,” Lockheed Aircraft Company,
Marietta, GA. (See also Goldstein, 1976.)
Lilley, G. M. (1974), “On the Noise from Jets,” AGARD Conference Proceedings, No. 131,
Chap. 13.
Lin, J. (2002), “Review of Research on Low-Profile Vortex Generators to Control
Boundary-Layer Separation,” Progress in Aerospace Sciences, Vol. 38, pp. 389–420.
Lin, J. C., Howard, F. G., Bushnell, D. M., and Selby, G. V. (1990), “Investigation of
Several Passive and Active Methods of Turbulent Flow Separation Control,” AIAA
Paper 90-1598.
Lin, J. C., Robinson, S. K., McGhee, R. J., and Valarezo, W. O. (1994), “Separation Control
on High-Lift Airfoils via Micro-Vortex Generators,” Journal of Aircraft, Vol. 31, No. 6,
pp. 1317–1323.
Linnick, M. N., and Fasel, H. F. (2005), “A High-Order Immersed Interface Method for
Simulating Unsteady Incompressible Flows on Irregular Domains,” Journal of
Computational Physics, Vol. 204, pp. 157–192.
Lions, J. L. (1971), Optimal Control of Systems Governed by Partial Differential Equations,
Springer, New York.
List, J., Byerley, A., Mclaughlin, T., and Van Dyken, B. (2003), “Using a Plasma Actuator
to Control Laminar Separation on a Linear Cascade Turbine Blade,” AIAA Paper
2003-1026.
Liu, J. T. C. (1974), “Developing Large-Scale Wavelike Eddies and the Near Jet Noise
Field,” Journal of Fluid Mechanics, Vol. 62, No. 3, pp. 437–464.
Liu, J. T. C. (1989), “Coherent Structures in Transisional and Turbulent Free Shear Flows,”
Annual Review of Fluid Mechanics, Vol. 21, pp. 285–315.
Liu, W. P., and Brodie, G. (1999), “A Demonstration of Active Turbulence Transition with
MEMS Sensors”, Abstract YC06.06, American Physical Society Centennial Meeting
Program, Atlanta, GA, March.
480 REFERENCES

Liu, C., Huang, C.-B., Zhu, Z., Jiang, F., Tung, S., Tai, Y.-C., and Ho, C.-M. (1999), “A
Micromachined Flow Shear-Stress Sensor Based on Thermal Transfer Principles,”
Journal of MEMS, Vol. 8, pp. 90–99.
Löfdahl, L., and Gad-el-Hak, M. (1999), “MEMS-Based Pressure and Shear Stress
Sensors for Turbulent Flows,” Measurement Science Technology, Vol. 10, No. 8,
pp. 665–686.
Löfdahl, L., Kälvesten, E., and Stemme, G. (1994), “Small Silicon Based Pressure
Transducers for Measurements in Turbulent Boundary Layers,” Experiments in Fluids,
Vol. 17, No. 1–2, pp. 24–31.
Long, T. A., and Petersen, R. A. (1992), “Controlled Interactions in a Forced Axisymmetric
Jet. Part 1. The Distortion of the Mean Flow,” Journal of Fluid Mechanics, Vol. 235,
pp. 37–55.
Lorber, P., McCormick, D., Anderson, T., Wake, B. and MacMartin, D., Pollack, M., Corke,
T., and Breuer, K. (2000), “Rotorcraft Retreating Blade Stall Control”, AIAA Paper
2000-2475.
Lou, H. C., Alvi, F. S., and Shih, C. (2006), “Active and Adaptive Control of Supersonic
Impinging Jets,” AIAA Journal, Vol. 44, No. 1, pp. 58–66.
Lubarsky, E., Shcherbik, D., Bibik, A., and Zinn, B. T. (2004), “Open Loop Control of
Severe Combustion Instabilities by Fuel Flow Modulation at Non-Resonant
Frequencies,” AIAA Paper 2004-0634.
Luchini, P., and Bottaro, A. (1998), “Görtler Vortices: a Backward-in-Time Approach to the
Receptivity Problem,” Journal of Fluid Mechanics, Vol. 363, pp. 1–23.
Luchtenburg, M., Tadmor, G., Lehmann, O., Noack, B. R., King, R., and Morzynski, M.
(2006), “Tuned POD Galerkin Models for Transient Feedback Regulation of the
Cylinder Wake,” AIAA Paper 2006-1407.
Lumley, J. L. (1967), “The Structure of Inhomogeneous Turbulent Flows,” Atmospheric
Turbulence and Radio Wave Propagation, edited by A. M. Yaglom and V. I. Tatarsky,
Nauka, Moscow, pp. 166–178.
Lumley, J. L. (1969), “Drag Reduction by Additives,” Annual Review of Fluid Mechanics,
Vol. 1, pp. 367–384.
Lumley, J. L. (1970), Stochastic Tools in Turbulence, Academic Press, New York.
Lumley, J., and Blossey, P. (1998), “Control of Turbulence,” Annual Review of Fluid
Mechanics, Vol. 30, pp. 311–327.
Lutz, T., and Wygnanski, I. (2005), “Controlling the Flow around a Swept Back Circular
Cylinder Using Periodic Excitation,” Bulletin of the American Physical Society, Nov.
Lysenko, V. I., and Maslov, A. A. (1984), “The Effect of Cooling on Supersonic Boundary-
Layer Stability,” Journal of Fluid Mechanics, Vol. 147, pp. 39–52.
Mabe, J. H., Calkins, F. T., Wesley, B., Woszildo, R., Taubert, L., Wygnanski, I. “Single
Dielectric Barrier Discharge Plasma Actuators for Improved Airfoil Performance,”
AIAA Journal of Aircraft. Vol. 46, No. 3, May–June 2009, pp. 847–845.
Mack, L. M. (1984), “Boundary Layer Linear Stability Theory,” AGARD-R-709,
pp. 3.1–3.62
MacMartin, D. G., Verma, A., Murray, R. M., and Paduano, J. D. (2001), “Active Control
of Integrated Inlet/Compression Systems: Initial Results,” ASME FEDSM2001–
18275.
Madou, M. (1997), Fundamentals of Microfabrication, CRC Press, New York.
Maestrello, L. (1977), “Statistical Properties of the Sound and Source Fields of an
Axisymmetric Jet,” AIAA Paper 1977-1267.
REFERENCES 481

Magill, J. C., and McManus, K. R. (2001), “Exploring the Feasibility of Pulsed Jet
Separation Control for Aircraft Configurations,” Journal of Aircraft, Vol. 38, No. 1,
pp. 48–56.
Mahmood, G., Hill, M., and Nelson, D. (2000), “Local Heat Transfer and Flow Structure
on and above a Dimpled Surface in a Channel,” ASME 2000-GT-230.
Maines, B., Smith, B., Baker, W., Saddoughi, S., and Shaw, L. (2006), “Effectiveness of an
Integrated Aerodynamic Wake and Structural Vibration Control System,” AIAA Paper
2006-439.
Maines, B., Smith, B., Baker, W., Saddoughi, S., and Shaw, L. (2007), “Effectiveness of an
Integrated Aerodynamic Wake and Structural Vibration Control System at Transonic
Speeds,” AIAA Paper 2007-878.
Maisel, M. D., Borgman, D. C., and Few, D. D. (1975), “NASA/Army XV-15 Tiltrotor
Research Aircraft Familiarization Document,” NASA TM X-62, 407.
Maisel, M., Laub, G., and McCroskey, W. J. (1986), “Aerodynamic Characteristics of
Two-Dimensional Wing Configurations at Angles of Attack Near-90°,” NASA
TM-88373.
Mankbadi, R. R. (1985), “On the Interaction between Fundamental and Subharmonic
Instability Waves in a Turbulent Round Jet,” Journal of Fluid Mechanics, Vol. 160,
pp. 385–419.
Mankbadi, R. R. (1992), “Dynamics and Control of Coherent Structure in Turbulent Jets,”
ASME Applied Mechchanics Review, Vol. 45, No. 6, pp. 219–248.
Manley, D., and von Klein, W. (2002), “Design and Development of a Super Short Takeoff
and Landing Transport Aircraft,” AIAA Paper 2002-6023.
Marasli, B., Champagne, F. H., and Wygnanski, I. (1992), “Effect of Travelling Waves on
the Growth of a Plane Turbulent Wake,” Journal of Fluid Mechanics, Vol. 235,
pp. 511–528.
Margalit, S., Greenblatt, D., Seifert, A., and Wygnanski, I. (2002), “Active Flow Control of
a Delta Wing at High Incidence using Segmented Piezoelectric Actuators,” AIAA
Paper 2002-3270.
Margalit, S., Greenblatt, D., Seifert, A., and Wygnanski, I. (2005), “Delta Wing Stall and
Roll Control using Segmented Piezoelectric Fluidic Actuators,” Journal of Aircraft,
Vol. 42, No. 3, pp. 698–709.
Margolin, L. G., and Rider, W. J. (2002), “A Rationale for Implicit Turbulence Modeling,”
International Journal for Numerical Methods in Fluids, Vol. 39, No. 9, pp.
821–841.
Marquardt, D. W. (1963), “An Algorithm for Least Squares Estimation of Nonlinear
Parameters,” Journal of Society of Industrial Applied Mathematics, Vol. 11, No. 2,
pp. 431–441.
Marquet, O., Sipp, D., and Jacquin, L. (2006), “Global Optimal Perturbations in a Separated
Flow over a Backward-Rounded-Step,” AIAA Paper 2006-2879.
Marsden, J. E., and Hoffman, M. J. (1998), Basic Complex Analysis, W. H. Freeman,
New York.
Mason, P. J. (1994), “Large-Eddy Simulation: a Critical Review of the Technique,”
Quarterly Journal of the Royal Meteorological Society, Vol. 120, pp. 1–26.
Massines, F., Gadri, R. Ben, Rabehi, A., Decomps, Ph., Segur, P., and Mayoux, Ch. (1998),
“Experimental and Theoretical Study of a Glow Discharge at Atmospheric Pressure
Controlled by Dielectric Barrier,” Journal of Applied Physics, Vol. 83, No. 6,
pp. 2950–2957.
482 REFERENCES

Mathew, J., Song, Q., Sankar, B., Sheplak, M., and Cattafesta, L. (2006), “Optimized
Design of Piezoelectric Flap Actuators for Active Flow Control,” AIAA Journal,
Vol. 44, No. 12, pp. 2919–2928.
Mattingly, J. D. (2006), Elements of Propulsion: Gas Turbines and Rockets, AIAA
Education Series, AIAA, New York.
Mattingly, J. D., Heiser, W. H., and Pratt, D. T. (2002), Aircraft Engine Design, AIAA
Education Series, AIAA, New York.
McCormick, D. C. (2000), “Boundary Layer Separation Control with Directed Synthetic
Jets,” AIAA Paper 2000-0519.
McCormick, D. C., Lozyniak, S., MacMartin, D. G., and Lorber, P. F. (2001), “Compact
High Power Boundary Layer Separation Control Actuation Development,” ASME
Paper FEDSM2001-18279.
McCroskey, W. J., McAlister, K. W., Carr, L. W., and Pucci, S. L. (1982), “An Experimental
Study of Dynamic Stall on Advanced Airfoil Sections. Vol. 2. Pressure and Force Data,”
NASA TM-84245, Vols. I–III.
McCullough, G. B., and Gault, D. E. (1949), “Boundary-Layer and Stalling Characteristics
of the NACA 64A0006 Airfoil Section,” NACA TN-1923.
McFadden, N. M., Rathert, G. A., Jr., and Bray, R. S. (1955), “The Effectiveness of
Wing Vortex Generators in Improving Maneuvering Characteristics of a Swept Wing
Airplane at Transonic Speeds,” NACA TN-3523.
McFarland, M. W. (Ed.) (2001), The Papers of Wilbur and Orville Wright, Including the
Chanute-Wright Papers, McGraw-Hill Companies.
McGowan, G., and Gopalarathnam, A. (2006), “Computational Study of a Circulation
Control Airfoil Using FLUENT,” Applications of Circulation Control Technology,
edited by R. D. Joslin and G. S. Jones, Vol. 214, AIAA Progress in Aeronautics and
Astronautics Series, Reston, VA, pp. 539–554.
McGowan, G., Gopalarathnam, A., Xiao, X., and Hassan, H. (2006), “Role of Turbulence
Modeling in Flow Prediction of Circulation Control Airfoils,” Applications of
Circulation Control Technology, edited by R. D. Joslin, and G. S. Jones, Vol. 214,
AIAA Progress in Aeronautics and Astronautics Series, Reston, VA, pp. 490–510.
McLaughlin, D. K., Morrison, G. L., and Troutt, T. R. (1975), “Experiments on the
Instability Waves in a Supersonic Jet and Their Acoustic Radiation,” Journal of Fluid
Mechanics, Vol. 69, No. 1, pp. 73–95.
McLean, J. D., Crouch, J. D., Stoner, R. C., Sakurai, S., Seidel, G. E., Feifel, W. M., and
Rush, H. M. (1999), “Study of the Application of Separation Control by Unsteady
Excitation to Civil Transport Aircraft,” NASA/CR-1999-209338.
McLean, J. D., Crouch, J. D., Stoner, R. C., Sakurai, S., Seidel, G. E., Feifel, W. M.,
Reimann, D., Pluim, J., and Bons, J. (2007), “Influence of Two-Dimensional vs.
Discrete Disturbances on Separating Low-pressure Turbine Boundary Layers,” AIAA
Paper 2007-525.
M’Closkey, R. T., King, J. M., Cortelezzi, L., and Karagozian, A. R. (2002), “The Actively
Controlled Jet in Crossflow,” Journal of Fluid Mechanics, Vol. 452, pp. 325–335.
McManus, K., and Magill, J. (1996), “Separation Control in Incompressible and
Compressible Flows using Pulsed Jets,” AIAA Paper 96-1948.
McManus, K. R., Vandsburger, U., and Bowman, C. T. (1990), “Combustor Performance
Enhancement through Direct Shear Layer Excitation,” Combustion and Flame, Vol. 82,
pp. 75–92.
REFERENCES 483

McManus, K. R., Poinsot, T., and Candel, S. (1993), “A Review of Active Control of
Combustion Instabilities,” Progress in Energy and Combustion Sciences, Vol. 19,
pp. 1–29.
McManus, K., Ducharme, A., Godley, C., and Magill, J. (1996), “Pulsed Jet Actuators for
Suppression of Flow Separation,” AIAA Paper 96-0442.
Meadowcroft, T., Dadone, L., and Sternfeld, H. (1992), “Advanced Concepts for Rotorcraft
Noise Reduction due to Blade–Vortex Interaction by Vortex Alleviation Devices,”
NASA CR-189626.
Mecham, M. (1996), “3M Thin Skin Tested by Airbus,” Aviation Week & Space Technology,
Vol. 145, No. 23.
Meitz, H. L., and Fasel, H. F. (2000), “A Compact-Difference Scheme for the Navier–Stokes
Equations in Vorticity-Velocity Formulation,” Journal of Computational Physics,
Vol. 157, No. 1, pp. 371–403.
Mellor, A. M. (1976), “Gas Turbine Engine Pollution,” Progress in Energy and Combustion
Science, Vol. 1, No. 2–3, pp. 111–133.
Menon, S. (1992), “Active Combustion Control in a Ramjet Using Large-Eddy Simulations,”
Combustion Science and Technology, Vol. 84, pp. 51–79.
Menon, S. (1995), “Secondary Fuel Injection Control of Combustion Instability in a
Ramjet,” Combustion Science and Technology, Vol. 100, pp. 385–393.
Menon, S. (2004), “CO Emission and Combustion Dynamics near Lean Blowout in Gas
Turbine Engines,” ASME Paper GT2004-53290.
Menon, S. (2005), “Acoustic–Vortex–Flame Interactions in Gas Turbine Combustors,”
Combustion Instabilities in Gas Turbine Engines: Operational Experience, Fundamental
Mechanisms, and Modeling, edited by T. Lieuwen and V. Yang, Vol. 210, Progress in
Astronautics and Aeronautics, pp. 277–314.
Menon, S., and Jou, W.-H. (1991), “Large-Eddy Simulations of Combustion Instability in
an Axisymmetric Ramjet Combustor,” Combustion Science and Technology, Vol. 75,
pp. 53–72.
Menon, S., and Yang, V. (1993), “Some Issues Concerning Active Control of Combustion
Instability in a Ramjet,” AIAA Paper 93-0116.
Menter, F. R. (1994), “Two-Equation Eddy-Viscosity Turbulence Models for Engineering
Applications,” AIAA Journal, Vol. 32, No. 8, pp. 1598–1605.
Menter, F. R., Kuntz, M., and Bender, R. (2003), “A Scale-Adaptive Simulation Model for
Turbulent Flow Predictions,” AIAA Paper 2003-0767.
Merchant, A., Braunsheidel, E., Kerrebrock, J. L., and Adamczyk, J. J. (2004), “Experimental
Investigation of a High Pressure Ratio Aspirated Fan Stage,” IGTI 2004, Vienna,
Austria, Paper GT2004-53679.
Merhaut, J. (1981), Theory of Electroacoustics, McGraw-Hill, New York.
Metcalfe, R. W., Orszag, S. A., Brachet, M. E., Menon, S., and Riley, J. J. (1987), “Secondary
Instability of a Temporally Growing Mixing Layer,” Journal of Fluid Mechanics,
Vol. 184, pp. 207–243.
Mezic, I., and Banaszuk, A. (2004), “Comparison of Systems with Complex Behavior,”
Physics D, Vol. 197, pp. 101–133.
Michalke, A. (1964), “On the Inviscid Instability of the Hyperbolic Tangent Velocity
Profile,” Journal of Fluid Mechanics, Vol. 19, No. 4, pp. 543–556.
Michalke, A. (1965), “Vortex Formation in a Free Boundary Layer According to Stability
Theory,” Journal of Fluid Mechanics, Vol. 22, No. 2, pp. 371–383.
484 REFERENCES

Michalke, A., and Fuchs, H. V. (1975), “On Turbulence and Noise of an Axisymmetric
Shear Flow,” Journal of Fluid Mechanics, Vol. 70, No. 1, pp. 179–205.
Michelassi, V., Wissink, J. G., Fröhlich, J., and Rodi, W. (2003), “Large-Eddy Simulation
of Flow Around Low-Pressure Turbine Blade with Incoming Wakes,” AIAA Journal,
Vol. 41, No. 11, pp. 2143–2156.
Milanovic, I. M., Zaman, K. B. M. Q., and Rumsey, C. L. (2005), “An Isolated Circular
Synthetic Jet in Crossflow at Low Momentum-Flux Ratio,” AIAA Paper 2005-1110.
Miller, D. N. (2004), “A Non-Intrusive Sensing Technique to Infer Optimal Flow Control
in Serpentine Inlets,” AIAA Paper 2004-1202.
Miller, D. N., and Addington, G. A. (2004), “Invited: Aerodynamic Flowfield Control
Technologies for Highly Integrated Propulsion/Airframe Flowpaths,” AIAA Paper
2004-2625.
Miller, D. N., and Catt, J. A. (1995), “Conceptual Development of Fixed-Geometry Nozzles
Using Fluidic Throat-Area Control,” AIAA Paper No. 95-2603.
Miller, D. N., and Smith, B. R. (2003), “CFD-Based Simulation of Inlet Unstart Phenomena:
Toward Supersonic Flow Control Techniques,” ASME FEDSM2003-45471.
Miller, D. N., Catt J. A., and Walker, S. H. (1997), “Extending Flow Control of Fixed
Nozzles Through Systematic Design: Introducing Assisted Reinjection,” ASME Paper
No. FEDSM97-3680.
Miller, D., Yagle, P., and Hamstra, J. (1999), “Fluidic Throat Skewing for Thrust Vectoring
in Fixed-Geometry Nozzles,” AIAA Paper 99-0365.
Miller, D., Yagle, P., Bender, E., and Vermeulen, P. (2001), “A Computational Investigation
of Pulsed Injection into a Confined Expanding Cross Flow,” AIAA Paper 2001-3026.
Milling, R. W. (1981), “Tollmien–Schlichting Wave Cancellation,” Physics of Fluids,
Vol. 24, No. 5, pp. 979–981.
Mittal, R., and Iaccarino, G. (2005), “Immersed Boundary Methods,” Annual Review of
Fluid Mechanics, Vol. 37, pp. 239–261.
Mittal, R., and Moin, P. (1997), “Suitability of Upwind-Biased Finite Difference Schemes
for Large Eddy Simulation of Turbulent Flows,” AIAA Journal, Vol. 35, No. 8,
pp. 1415–1417.
Mittal, R., Rampunggoon, P., and Udaykumar, H. S. (2001), “Interaction of a Synthetic Jet
with a Flat Plate Boundary Layer,” AIAA Paper 2001-2773.
Mohd-Yusof, J. (1997), “Combined Immersed-Boundary/B-Spline Methods for Simulations
of Flow in Complex Geometries,” Annual Research Briefs, Center for Turbulence
Research, Stanford, CA, pp. 317–328.
Moin, P., and Bewley, T. (1994), “Feedback Control of Turbulence,” Applied Mechanics
Reviews, Vol. 47, pp. S3-S13.
Moler, C., and Loan, C. V. (2003), “Nineteen Dubious Ways to Compute the Exponential
of a Matrix, Twenty-Five Years Later,” SIAM Review, Vol. 45, No. 1, pp. 1–46.
Mongia, H. C., Santoro, R., Menon, S., Gutmark, E., and Gore, J. (2001), “Combustion
Research Needs for Helping Development of Next-Generation Combustors,” AIAA
Paper 2001-3853.
Mongia, H. C., Held, T. J., Hsaiao, G. C., and Pandalai, R. P. (2003), “Challenges and
Progress in Controlling Dynamics in Gas Turbine Combustors,” Journal of Propulsion
and Power, Vol. 19, No. 5, pp. 822–829.
Monkewitz, P. A. (1989), “Feedback Control of Global Oscillations in Fluid Systems,”
AIAA Paper 89-0991.
Moore, C. J. (1977), “The Role of Shear-layer Instability Waves in Jet Exhaust Noise,”
Journal of Fluid Mechanics, Vol. 80, No. 2, pp. 321–367.
REFERENCES 485

Moore, B. C. (1981), “Principal Component Analysis in Linear Systems: Controllability,


Observability, and Model Reduction,” IEEE Transactions on Automatic Control,
Vol. 26, No. 1, pp. 17–32.
Moore, F. K., and Greitzer, E. M. (1986), “A Theory of Post-Stall Transients in Axial Flow
Compression Systems Part I: Development of Equations,” ASME Journal of Engineer-
ing for Gas Turbines and Power, Vol. 108, pp. 68–76.
Moore, D. W., and Saffman, P. G. (1973), “Axial Flow in Laminar Trailing Vortices,”
Proceedings of the Royal Society London Series A, Vol. 333, pp. 491–508.
Moore, F. K., Getizer E. M., Greitzer, E. M., and Moore, F. K. (1986), “A Theory of
Post-Stall Transients in Axial Compression Systems Parts I & II,” Transactions of the
ASME Journal of Engineering for Gas Turbines and Power, Vol. 108, pp. 231–239.
Moreau, E. (2007), “Airflow Control by Non-Thermal Plasma Actuators,” Journal of
Physics D: Applied Physics, Vol. 40, pp. 605–636.
Morfey, C. L. (1973), “Amplification of Aerodynamic Noise by Convected Flow
Inhomogeneities,” Journal of Sound and Vibration, Vol. 31, pp. 391–397.
Morgan, P. E., Rizzetta, D. P., and Visbal, M. R. (2005a), “Large-Eddy Simulation of Flow
over a Wall-Mounted Hump,” AIAA Paper 2005-0484.
Morgan, P. E., Rizzetta, D. P., and Visbal, M. R. (2005b), “Large-Eddy Simulation of
Separation Control for Flow over a Wall-Mounted Hump,” AIAA Paper 2005-5017.
Morgan, P. E., Rizzetta, D. P., and Visbal, M. R. (2006), “High-Order Numerical Simulation
of Turbulent Flow over a Wall-Mounted Hump,” AIAA Journal, Vol. 44, No. 2,
pp. 239–251.
Morgans, A. S., and Dowling, A. P. (2007), “Model-Based Control of Combustion
Instabilites,” Journal of Sound and Vibration, Vol. 299, pp. 261–282.
Morgans, A. S., and Stow, S. R. (2007), “Model-Based Control of Combustion Instabilities
in Annular Combustors,” Combustion and Flame, Vol. 150, pp. 380–399.
Morkovin, M. V. (1969), “On the Many Faces of Transition,” Viscous Drag Reduction,
edited by C. S. Wells, Plenum Press, New York, pp. 1–32.
Morosanov, I. S. (1957), “Method of Extremum Control,” Automatic and Remote Control,
Vol. 18, pp. 1077–1092.
Morse, P. M., and Feshbach, H. (1953), Methods of Theoretical Physics, Parts I and II,
McGraw-Hill, New York.
Morzynski, M., and Afanasiev, K. (1996), “Lösung von Grossen Unsymmetrischen
Eigenwertproblemen bei der Stabilitätsanalyse auf der Grundlage der 2-D Navier–
Stokes-Gleichungen,” Z. Angew. Math. Mech., Vol. 77, No. 2, pp. 627–628.
Morzynski, M., Afanasiev, K., and Thiele, F. (1999), “Solution of the Eigenvalue Problems
Resulting from Global Nonparallel Flow Stability Analysis,” Computational Methods
in Applied Mechanical Engineering, Vol. 169, pp. 161–176.
Mounts, J., and Barber, T. (1992), “Numerical Analysis of Shock-Induced Separation
Alleviation Using Vortex Generators,” AIAA Paper 92-0751.
Mueller T. J. (1999), “Aerodynamic Measurements at Low Reynolds Numbers for Fixed
Wing Micro-Air Vehicles,” presented at the RTO AVT/VKI Special Course on
Development and Operation of UAVs for Military and Civil Applications, VKI,
Belgium.
Muller, M. O., Bernal, L. P., Miska, P. K., Washabaugh, P. D., and Chou, T. K. A. (2001),
“Flow Structure and Performance of Axisymmetric Synthetic Jets,” AIAA Paper 2001-
1008.
Munro, S. E., and Ahuja, K. K. (2003), “Aeroacoustics of a High Aspect-Ratio Jet,” AIAA
Paper 2003-3323.
486 REFERENCES

Munro, S. E., Ahuja, K. K., and Englar, R. J. (2001), “Noise Reduction through Circulation
Control,” AIAA Paper 2001-666.
Murugappan, S., Acharya, S., Allgood, D. C., Park, S., Annaswamy, A. M., and Ghoniem,
A. F. (2003), “Optimal Control of a Swirl Stabilized Spray Combustor using System
Identification Approach,” Combustion Science and Technology, Vol. 175, pp. 55–81.
Muruganandam, T. M., Nair, S., Scarborough, D., Neumeier, Y., Jagoda, J., Lieuwen, T.,
Seitzman, J., and Zinn, B. T. (2005), “Active Control of Lean Blowout for Turbine
Engine Combustors,” Journal of Propulsion and Power, Vol. 21, No. 5, pp. 807–814.
Musiyenko, V. P. (1993), “Experimental Study of the Flow around Localized Deepenings,”
Bionics, No. 26, pp. 31–34.
Musquere, A. (2003), “Vaincre les Defauts des Convertibles,” Air et Cosmos, No. 1901,
p. 34.
Mustafa, D., and Glover, K. (1990), Minimum Entropy H• Control, Springer-Verlag,
Berlin.
Nabi, A., Neumeier, Y., Sattinger, S., and Zinn, B. (2000), “Sub-Scale Demonstration of the
Active Feedback Control of Gas-Turbine Combustion Instabilities,” Journal of
Engineering for Gas Turbines and Power, Vol. 122, pp. 262–268.
Nagel, A., Levy, D. E., and Shepshelovich, M. (2006), “Conceptual Aerodynamic Evaluation
of MINI/MICRO UAV,” AIAA Paper 2006-1261.
Nagib, H., Greenblatt, D., Kiedaisch, J., Wygnanski, I., and Hassan, A. (2001), “Effective
Flow Control for Rotorcraft Applications at Flight Mach Numbers,” AIAA Paper 2001-
2974.
Nagib, H. M., Kiedaisch, J. W., Wygnanski, I. J., Stalker, A. D., Wood, T., and McVeigh,
M. A. (2004), “First In-Flight Full-Scale Application of Active Flow Control: The
XV-15 Tiltrotor Download Reduction,” NATO RTO-MP-AVT-111-P-29, Prague,
Czech Republic, October.
Nagib, H., Kiedaisch, J., Reinhard, P., and Demanett, B. (2006), “Control Techniques for
Flows with Large Separated Regions: a New Look at Scaling Parameters,” AIAA Paper
2006-2857.
Nagib, H., Kiedaisch, J., Greenblatt, D., Wygnanski, I., and Hassan, A. (2007a), “Flow
Control for Rotorcraft Applications at Flight Mach Numbers,” IUTAM Symposium on
Unsteady Separated Flows and their Control, Cyprus.
Nagib, H., Kiedaisch, J., Reinhard, P., and Demanett, B. (2007b), “Active Flow Control
for High Lift Airfoils: Separation Versus Circulation Control,” AIAA Paper 2007-
1119.
Naim, A., Greenblatt, D., Seifert A., and Wygnanski, I. (2002), “Active Control of a Circular
Cylinder Flow at Transitional Reynolds Numbers,” AIAA Paper 2002-3070.
Naim, A., Greenblatt, D., Seifert, A., and Wygnanski, I. (2007), “Active Control of a
Circular Cylinder Flow at Transitional Reynolds Numbers,” Flow, Turbulence and
Combustion—Special Issue on Air-Jet Actuators and Their Use for Flow Control,
Vol. 78, pp. 383–407.
Nair, S., Thiruchengode, M., Olsen, R., Meyer, A., Seitzman, J., Zinn, B. T., Lieuwen, T.,
Held, T., and Mongia, H. (2004), “Lean Blowout Detection in a Single Nozzle, Swirl
Cup Combustor,” AIAA Paper 2004-0138.
Nair, S., and Lieuwen, T. (2005), “Acoustic Detection of Blowout in Premixed Flames,”
Journal of Propulsion and Power, Vol. 21, No. 1, pp. 32–39.
Nair, S., and Lieuwen, T. (2007), “Near-Blowoff Dynamics of a Bluff-Body Stabilized
Flame,” Journal of Propulsion and Power, Vol. 23, No. 2, pp. 421–429.
REFERENCES 487

NASA (2007), “Advanced Concept Studies for Subsonic and Supersonic Commercial
Transports Entering Service in the 2030–35 Period.” Research Announcement,
Pre-proposal Conference, Washington, DC, 29 November (http://www.aeronautics.
nasa.gov/pdf/n3_draft_summary_11_29_07.pdf).
Naughton, J. W., and Sheplak, M. (2002), “Modern Developments in Shear Stress
Measurement,” Progress in Aerospace Sciences, Vol. 38, No. 6–7, pp. 515–570.
Naughton, J. W., Viken, S., and Greenblatt, D. (2006), “Skin-Friction Measurements on the
NASA Hump Model,” AIAA Journal, Vol. 44, No. 6, pp. 1255–1265.
Naveh, T., Seifert, A., Tumin, A., and Wygnanski, I. (1998), “Sweep Effect on the Parameters
Governing the Control of Separation by Periodic Excitation,” Journal of Aircraft,
Vol. 35, No. 3, pp. 510–512.
Nedungadi, A., Barber, T. J., Nishimura, M., and Kudo, T. (2001), “The Effects of Pulsed
Blowing on Jet Mixing and Noise Generation,” AIAA Paper 2001-0665.
Nelson, P. A. (2000), “Active Techniques and Their Potential for Application in
Aeroacoustics,” AIAA Paper 2000-2100.
Nelson, G. M., and Lakany, H. (2007), “An Investigation Into the Application of Fuzzy
Logic Control to Industrial Gas Turbines,” ASME Journal of Engineering for Gas
Turbines and Power, Vol. 129, No. 4, pp. 1138–1142.
Neuburger, D., and Wygnanski, I. (1987), “The Use of a Vibrating Ribbon to Delay
Separation on Two-Dimensional Airfoils: Some Preliminary Observations,” Presented
at the Workshop on Unsteady Separated Flow, Air Force Academy, July.
Neuendorf, R., and Wygnanski, I. (1999), “On a Turbulent Wall Jet Flowing over a Circular
Cylinder,” Journal of Fluid Mechanics, Vol. 381, pp. 1–25.
Neuendorf, R., Lourenco, L., and Wygnanski, I. (2004), “On Large Streamwise Structures
in a Wall Jet Flowing over a Circular Cylinder,” Physics of Fluids, Vol. 16, No. 7,
pp. 2158–2169.
Neumann, J., and Wengle, H. (2001), “Active Control of Turbulent Separated Flows Using
Large-Eddy Simulation,” Direct and Large-Eddy Simulation IV, edited by B. J. Geurts,
R. Friedrich, and O. Metais, Kluwer Academic Publishers, Dordrecht, pp. 427–434.
Neumann, J., and Wengle, H. (2003), “DNS and LES of Passively Controlled Turbulent
Backward-Facing Step Flow,” Flow, Turbulence, and Combustion, Vol. 71, pp. 297–310.
Neumeier, Y., and Zinn, B. T. (1996a), “Active Control of Combustion Instabilities with
Real Time Observation of Unstable Combustor Modes,” AIAA Paper 1996-0758.
Neumeier, Y., and Zinn, B. T. (1996b), “Experimental Demonstration of Active Control of
Combustion Instabilities Using Real Time Modes Observation and Secondary Fuel
Injection,” Proceedings of the Combustion Institute, Vol. 26, pp. 2811–2818.
Neumeier, Y., and Zinn, B. T. (1996c), “Investigation of the Open Loop Performance of an
Active Control System Utilizing a Fuel Injector Actuator,” AIAA Paper 1996-2757.
Neumeier, Y., Arbel, A., Nabi, A., Vertzberger, M., and Zinn, B. T. (1997), “Open-Loop
Performance of a Fast-Response, Actively Controlled Fuel Injector Actuator,” Journal
of Propulsion and Power, Vol. 13, No. 6, pp. 705–713.
Ng, T. T., and Bradley, T. A. (1988), “Effect of Multifrequency Forcing on the Near-Field
Development of a Jet,” AIAA Journal, Vol. 26, No. 10, pp. 1201–1207.
Ng, W. F., and Burdisso, R. A. (2001), “Active Acoustic and Flow Control for
Aeropropulsion,” AIAA Paper 2001-0220.
Nguyen, K., Betzina, M., and Kitaplioglu, C. (2000), “Full-Scale Demonstration of Higher
Harmonic Control for Noise and Vibration Reduction on the XV-15 Rotor”, presented
at the American Helicopter Society 56th Annual Forum, Virginia Beach, VA, May.
488 REFERENCES

Nickol, C. L. (2008), “Silent Aircraft Initiative Concept Risk Assessment,” NASA TM


2008–215112.
Nishri, B. (1995), On the Dominant Mechanism of Active Control of Flow Separation,
Ph.D. Thesis, Tel-Aviv Univ., Israel.
Nishri, B., and Wygnanski, I. (1998), “Effects of Periodic Excitation on Turbulent Flow
Separation from a Flap,” AIAA Journal, Vol. 36, No. 4, pp. 547–556.
Noack, B., Afanasiev, K., Morzynśki, M., Tadmor, G., and Thiele, F. (2003), “A Hierarchy
of Low-Dimensional Models for the Transient and Post-Transient Cylinder Wake,”
Journal of Fluid Mechanics, Vol. 497, pp. 335–363.
Noack, B. R., Tadmor, G., and Morzynski, M. (2004), “Low-Dimensional Models
for Feedback Flow Control. Part I: Empirical Galerkin Models,” AIAA Paper 2004-
2408.
Nokhimovitch, M., Benima, R., Aharonov, H., and Seifert, A. (2002), “Closed-Loop
Control of a Ball Position on a Beam using Piezoelectric Fluidic Jets,” Proceedings of
the 42nd ISR Aero Meeting, Israel.
North, W. J. (1955), “Summary Evaluation of Toothed-Nozzle Attachments as a Jet-Noise-
Suppression Device,” NACA-TN-3516.
Nuber, R. J., and Needham, J. R., Jr. (1948), “Exploratory Wind-Tunnel Investigation of the
Effectiveness of Area Suction in Eliminating Leading-Edge Separation over an NACA
641A212 Airfoil,” NACA TN-1741.
Nygaard, K. J., and Glezer, A. (1991), “Evolution of Streamwise Vortices and Generation
of Small-Scale Motion in a Plane Mixing,” Journal of Fluid Mechanics, Vol. 231,
pp. 257–301.
Nygaard, K. J., and Glezer, A. (1994), “The Effect of Phase Variations and Cross-Shear on
Vortical Structures in a Plane Mixing Layer,” Journal of Fluid Mechanics, Vol. 276,
pp. 21–59.
Oates, G. C. (1989), Aircraft Propulsion Systems Technology and Design, AIAA Education
Series, New York.
Obinata, G., and Anderson, B. D. O. (2000), Model Reduction for Control System Design,
Springer-Verlag, Berlin.
Ogata, K. (2001), Modern Control Engineering, 4th ed., Prentice-Hall, Upper Saddle
River, NJ.
Olson, D. H., Reimann, D., Bloxham, M., and Bons, J. P. (2005), “The Effect of Elevated
Freestream Turbulence on Separation Control with Vortex-Generating Jets,” AIAA
Paper 2005-1114.
Opila, D., Annaswamy, A. M., Krol, W. P., and Raghu, S. (2004), “Biomimetic Reduction
of Wake Deficit Using Tail Articulation at Low Reynolds Number,” IEEE Journal of
Oceanic Engineering, Vol. 29, No. 3, pp. 766–776.
Orlov, D. M., and Corke, T. C. (2005), “Numerical Simulation of Aerodynamic Plasma
Actuator Effects,” AIAA Paper 2005-1083.
Oster, D., and Wygnanski, I. (1982), “The Forced Mixing Layer Between Parallel Streams,”
Journal of Fluid Mechanics, Vol. 123, pp. 91–130.
Oster, D., Wygnanski, I., Dziomba, B., and Fiedler, H. (1978), “The Effect of Initial
Conditions on the Two-Dimensional, Turbulent Mixing Layer,” Structure and
Mechanics of Turbulence, edited by H. Fiedler, Vol. 75, Lecture Notes in Physics,
Springer, Berlin, pp. 48–64.
Ostrovskii, I. I. (1957), “Extremum Regulation,” Automatic and Remote Control, Vol. 18,
pp. 900–907.
REFERENCES 489

Owens, L. R, Allan, B. G., and Gorton, S. A. (2006), “Boundary-Layer Ingesting Inlet Flow
Control,” AIAA Paper 2006-0839.
Pack, L. G., and Seifert, A. (1999), “Periodic Excitation for Jet Vectoring and Enhanced
Spreading,” AIAA Paper 1999-0672.
Pack, L. G., and Seifert, A. (2001a), “Multiple Mode Actuation of a Turbulent Jet,” AIAA
Paper 2001-0735.
Pack, L. G., and Seifert, A. (2001b), “Periodic Excitation for Jet Vectoring and Enhanced
Spreading,”Journal of Aircraft, Vol. 38, No. 3, pp. 486–495.
Pack-Melton, L. G., Yao, C. S., and Seifert, A. (2004), “Application of Excitation from
Multiple Locations on a Simplified High-Lift System,” AIAA Paper 2004-2324.
Pack-Melton, L. G., Schaeffler, N., Yao, C. S., and Seifert, A. (2005), “Active Control of
Flow Separation from the Slat Shoulder of a Supercritical Airfoil,” Journal of Aircraft,
Vol. 42, No. 5, pp. 1142–1149.
Pack-Melton, L. G., Yao, C. S., and Seifert, A. (2006), “Active Control of Flow Separation
from the Flap of a Supercritical Airfoil,” AIAA Journal, Vol. 44, No. 1, pp. 34–41.
Pack-Melton, L. G., Schaeffler, N., and Lin, J. (2007), “High-Lift System for a Supercritical
Airfoil: Simplified by Active Flow Control,” AIAA Paper 2007-0707.
Padmanabhan, K. T., Bowman, C. T., and Powell, J. D. (1995), “An Adaptive Optimal
Combustion Control Strategy,” Combustion and Flame, Vol. 100, pp. 101–110.
Paduano, J. D., Greitzer, E. M., and Epstein, A. H. (2001), “Compression System Stability
and Active Control,” Annual Review of Fluid Mechanics, Vol. 33, pp. 491–517.
Pal, D., and Sinha, K. (1997), “Controlling an Unsteady Separating Boundary Layer on a
Cylinder with an Active Compliant Wall,” AIAA Paper 97-0212.
Palaghita, T. I., and Seitzman, J. (2004), “Control of Temperature Non-Uniformity Based
on Line-of-Sight Absorption,” AIAA Paper 2004-4163.
Palaghita, T. I., and Seitzman, J. (2005), “Pattern Factor Sensing and Control Based on
Diode Laser Absorption,” AIAA Paper 2005-3578.
Pandalai, R. P., and Mongia, H. C. (1998), “Combustion Instability Characteristics
of Industrial Engine Dry Low Emission Combustion Systems,” AIAA Paper 1998-3379.
Pankhurst, R. C., and Gregory, B. A. (1948), “Power Requirements for Distributed Suction
for Increasing Maximum Lift,” Aeronautical Research Council CP, No. 82.
Pankhurst, R. C., Raymer, W. G., and Devereaux, A. N. (1948), “Wind-Tunnel Tests of the
Stalling Properties of an 8 percent Thick Symmetrical Section with Nose Suction
Through a Porous Surface,” Aeronautical Research Council R. & M., No. 2666.
Parekh, D. E., and Glezer, A. (2000), “AVIA—Adaptive Virtual Aerosurface,” AIAA Paper
2000-2474.
Parekh, D. E., Kibens, V., Glezer, A., Wiltse, J. M., and Smith D. M. (1996), “Innovative
Jet Flow Control—Mixing Enhancement Experiments,” AIAA Paper 96-0308.
Park, H. M., and Lee, M. W. (1998), “An Efficient Method of Solving the Navier–Stokes
Equations for Flow Control,” International Journal for Numerical Methods in
Engineering, Vol. 41, pp. 1133–1151.
Park, Y. W., Lee, S.-G., Lee, D.-H., and Hong, S. (2001), “Stall Control with Local Surface
Buzzing on a NACA 0012 Airfoil,” AIAA Journal, Vol. 39, No. 7, pp. 1400–1402.
Park, S. H., Yu, Y. H., and Byun, D. Y. (2007), “RANS Simulation of a Synthetic Jet in
Quiescent Air,” AIAA Paper 2007-1131.
Paschen, F. (1889), “Ueber die zum Funkenübergang in Luft, Wasserstoff und Kohlensäure
bei verschiedenen Drucken erforderliche Potentialdifferenz,” Annalen der Physik,
Vol. 273, No. 5, pp. 69–75.
490 REFERENCES

Paschereit, C. O., and Gutmark, E. (2002), “Proportional Control of Combustion Instabilities


in a Simulated Gas-Turbine Combustor,” Journal of Propulsion and Power, Vol. 18,
No. 6, pp. 1298–1304.
Paschereit, C. O., Wygnanski I., and Fiedler, H. E. (1995), “Experimental Investigation of
Subharmonic Resonance in an Axisymmetric Jet,” Journal of Fluid Mechanics,
Vol. 283, pp. 365–407.
Paschereit, C. O., Gutmark, E., and Weisenstein, W. (1998), “Structure and Control of
Thermoacoustic Instabilities in a Gas Turbine Combustor,” Combustion Science and
Technology, Vol. 138, pp. 213–232.
Paschereit, C. O., Gutmark, E., and Weisenstein, W. (1999), “Coherent Structures in
Swirling Flows and their Role in Acoustic Combustion Control,” Physics of Fluids,
Vol. 11, pp. 2667–2678.
Patel, V. C., and Sotiropoulos, F. (1997), “Longitudinal Curvature Effects in Turbulent
Boundary Layers,” Progress in Aerospace Sciences, Vol. 33, Issue 1–2, pp. 1–70.
Paterson, E. G., and Baker, W. J. (2004), “Simulation of Steady Circulation Control for
Marine-Vehicle Control Surfaces,” AIAA Paper 2004-0748.
Paterson, E. G., and Baker, W. J. (2006), “RANS and Detached-Eddy Simulation of the
NCCR Airfoil,” Applications of Circulation Control Technology, edited by R. D. Joslin
and G. S. Jones, Vol. 214, AIAA Progress in Aeronautics and Astronautics Series,
Reston, VA, pp. 421–444.
Pegg, R. J., Hosier, R. N., Balcerak, J. C., and Johnson, H. K. (1975), “Design and
Preliminary Tests of a Blade Tip Air Mass Injection System for Vortex Modification
and Possible Noise Reduction on A Full-Scale Helicopter Rotor,” NASA TM X-3314.
Peskin, C. S. (1972), “Flow Patterns Around Heart Valves: A Numerical Method,” Journal
of Computational Physics, Vol. 10, pp. 252–271.
Peskin, C. S. (1981), “The Fluid Dynamics of Heart Valves: Experimental, Theoretical and
Computational Methods,” Annual Review of Fluid Mechanics, Vol. 14, pp. 235–259.
Peterka, J. A., and Richardson, P. D. (1969), “Effects of Sound on Separated Flow,” Journal
of Fluid Mechanics, Vol. 37, pp. 265–287.
Peterlin, A. (1961), “Streaming Birefringence of Soft Linear Macromolecules with Finite
Chain Length,” Polymer 2, pp. 257–264.
Phillips, E. H. (2003), “Darpa Dumps Drag,” Aviation Week and Space Technology,
Vol. 159, No. 3, p. 30.
Piomelli, U., and Balaras, E. (2002), “Wall-Layer Models for Large Eddy Simulations,”
Annual Review of Fluid Mechanics, Vol. 34, pp. 349–374.
Piomelli, U., Ferziger, J., Moin, P., and Kim, J. (1989), “New Approximate Boundary
Conditions for Large Eddy Simulations of Wall-Bounded Flows,” Physics of Fluids A,
Vol. 1, No. 6, pp. 1061–1068.
Piomelli, U., Balaras, E., Squires, K. D., and Spalart, P. R. (2002), “Interaction of the Inner
and Outer Layers in Large-Eddy Simulations with Wall-Layer Models,” Engineering
Turbulence Modelling and Experiments 5, edited by W. Rodi and N. Fueyo, Elsevier
Science, Amsterdam, pp. 307–316.
Pitsch, H. (2006), “Large-Eddy Simulation of Turbulent Combustion,” Annual Review of
Fluid Mechanics, Vol. 38, pp. 453–482.
Plas, A. P., Sargeant, M. A., Madani, V., Crichton, D., Greitzer, E. M., Hynes, T. P., and Hall,
C. A. (2007), “Performance of BLI Propulsion System,” AIAA Paper 2007-0450.
Poinsot, T., and Veynante, D. (2005), Theoretical and Numerical Combustion, 2nd ed.,
R. T. Edwards, Philadelphia, PA.
REFERENCES 491

Poinsot, T., Trouve, R. F., Veynante, D. P., Candel, S. M., and Esposito, E. J. (1987), “Vortex
Driven Acoustically Coupled Combustion Instabilities,” Journal of Fluid Mechanics,
Vol. 177, pp. 262–292.
Poinsot, T., Veynante, D., Bourienne, F., Candel, S., Esposito, E., and Surjet, J. (1988),
“Initiation and Suppression of Combustion Instabilities by Active Control,” Proceedings
of the Combustion Institute, Vol. 22, pp. 1363–1370.
Poisson-Quinton, Ph. (1948), “Recherches théoriques et expérimentales sur le contrôl de
couche limits,” 7th Congress of Applied Mechanics, London.
Poisson-Quinton, Ph., and Lepage, L. (1961), “Survey of French Research on Control of
Boundary Layer and Circulation,” Boundary Layer and Flow Control, Vol. 1, edited by
G.V. Lachmann, Pergamon Press, New York, pp. 21–73.
Polacsek, C., and Desbois-Lavergne, F. (2003), “Fan Interaction Noise Reduction using a
Wake Generator: Experiments and Computational Aeroacoustics,” Journal of Sound
and Vibration, Vol. 265, No. 4, pp. 725–743.
Poll, D. I. A. (1985), “Some Observations of the Transition Process on the Windward
Face of a Long Yawed Cylinder,” Journal of Fluid Mechanics, Vol. 150, pp.
329–356.
Polychroniadis, M. (1990), “Generalized Higher Harmonic Control—Ten Years of
Aerospatiale Experience,” Paper No. III.7.2, Sixteenth European Rotorcraft Forum,
Glasgow, Scotland, Sept.
Pope, S. B. (2000), Turbulent Flows, Cambridge University Press, Cambridge, UK.
Post, M. L., and Corke, T. C. (2004), “Separation Control on High Angle of Attack Airfoil
Using Plasma Actuators,” AIAA Journal, Vol. 42, No. 11, pp. 2177–2182.
Post, M. L., and Corke, T. C. (2006), “Separation Control Using Plasma Actuators: Dynamic
Stall Vortex Control on Oscillating Airfoil,” AIAA Journal, Vol. 44, No. 12,
pp. 3125–3135.
Postl, D. (2005), Numerical Investigation of Laminar Separation Control Using Vortex
Generator Jets, Ph.D. Thesis, University of Arizona, Tucson, AZ.
Postl, D., Gross, A., and Fasel, H. F. (2003), “Numerical Investigation of Low-Pressure
Turbine Blade Separation Control,” AIAA Paper 2003-0614.
Postl, D., Gross, A., and Fasel, H. F. (2004), “Numerical Investigation of Active Flow
Control for Low-Pressure Turbine Blade Separation,” AIAA Paper 2004-0750.
Postl, D., and Fasel, H. F. (2006), “Direct Numerical Simulation of Turbulent Flow
Separation from a Wall-Mounted Hump,” AIAA Journal, Vol. 44, No. 2, pp.
263–272.
Praisner, T. J., and Clark, J. P. (2004), “Predicting Transition in Turbomachinery, Part 1—A
Review and New Model Development,” ASME Paper No. GT-2004-54108.
Praisner, T. J., Grover, E. A., Rice, M. J., and Clark, J. P. (2004), “Predicting Transition in
Turbomachinery, Part 2—Model Validation and Benchmarking,” ASME Paper
GT-2004-54109.
Prakash, S., Nair, S., Muruganandam, T. M., Neumeier, Y., Lieuwen, T., Seitzman, J., and
Zinn, B. T. (2005), “Acoustic Sensing and Mitigation of Lean Blow Out in Premixed
Flames,” AIAA Paper 2005-1420.
Pralits, J. O., and Hanifi, A. (2003), “Optimization of Steady Suction for Disturbance
Control on Infinite Swept Wings,” Physics of Fluids, Vol. 15, No. 9, pp. 2756–2772.
Pralits, J. O., Airiau, C., Hanifi, A., and Henningson, D. S. (2000), “Sensitivity Analysis
using Adjoint Parabolized Stability Equations for Compressible Flows,” Flow,
Turbulence, Combustion, Vol. 65, p. 321.
492 REFERENCES

Pralits, J. O., Hanifi, A., and Henningson, D. S. (2002), “Adjoint-Based Optimization of


Steady Suction for Disturbance Control in Incompressible Flows,” Journal of Fluid
Mechanics, Vol. 467, pp. 129–161.
Prandtl, L. (1904), “Uber Flussigkeitsbewegug bei sehr kleiner Reibung,” Third International
Congress of Mathematicians, Heidelberg, pp. 484–491. (Also, Vier Abhandlungen zur
Hydro-dynamik und Aerodynamik”, pp. 1–8, Gottingen, 1927). (English Translation:
Motion of Fluids with Very Little Viscosity, NACA TM-452, 1928.)
Prandtl, L., and Tietjens, O. (1929, 1934), Hydro- und Aeromechanik (based on Prandtl’s
lectures), Vols. I and II, Berlin, 1929 and 1931. (English Translation L. Rosenhead
(Vol. I) and J.P. den Hartog (Vol. II), New York, 1934.)
Prasad, S., Gallas, Q., Horowitz, S., Homeijer, B., Sankar, B., Cattafesta, L., and Sheplak,
M. (2006), “An Analytical Electroacoustic Model of a Piezoelectric Composite Circular
Plate,” AIAA Journal, Vol. 44, No. 10, pp. 2311–2318.
Prasanth, R. K., Annaswamy, A. M., Harhout, J. P., and Ghoniem, A. F. (2002), “When Do
Open-Loop Strategies for Combustion Control Work?” Journal of Propulsion and
Power, Vol. 18, No. 3, pp. 658–668.
Raghavan, V., Chopra, I., and Pai, S. (1988), “Circulation Control Airfoils in Unsteady
Flow,” Journal of the American Helicopter Society, Vol. 33, pp. 28–37.
Raju, R., Mittal, R., and Cattafesta, L. N., III (2003), “Towards Physics Based Strategies
for Separation Control over an Airfoil Using Synthetic Jets,” AIAA Paper 2007-1421.
Raman, G. (1998), “Advances in Understanding Supersonic Jet Screech: Review and
Perspective,” Progress in Aerospace Sciences, Vol. 34, pp. 45–106.
Raman, G., and Cain, A. B. (2002), “Innovative Actuators for Active Flow and Noise
Control,” Proceedings of IMechE, Part G, Journal of Aerospace Engineering, Vol. 216,
pp. 303–323.
Raman, G., and Cornelius, D. (1994), “Jet Mixing Control Using Excitation from Miniature
Oscillating Jets,” AIAA Journal, Vol. 33, No. 2, pp. 365–368.
Raman, G., and Kibens, V. (2001), “Active Flow Control using Integrated Powered
Resonance Tube Actuators,” AIAA Paper 2001-3024.
Raman, G., and Raghu, S. (2000), “Miniature Fluidic Oscillators for Flow Control and
Noise Control,” AIAA Paper 2000-2554.
Raman, G., and Rice, E. (1991), “Axisymmetric Jet Forced by Fundamental and
Subharmonic Tones,” AIAA Journal, Vol. 29, No. 7, pp. 1114–1122.
Raman, G., Mills, A., Othman, S., and Kibens, V. (2001), “Development of Powered
Resonance Tube Actuators for Active Flow Control,” ASME FEDSM 2001-18273.
Raman, G., Khanafseh, S., and Cain, A. (2002), “Development of High Bandwidth
Actuators for Aeroacoustic Control,” AIAA Paper 2002-0664.
Raman, G., Sarpotdar, S., Tassy, J., Cain, A., and Kerschen, E. (2004a), “An Overview of
the Development of High Bandwidth Powered Resonance Tube Actuators: Experiments
and Simulations,” AIAA Paper 2004-2856.
Raman, G., Khanafseh, S., Cain, A. B., and Kerchen, E. (2004b), “Development of High
Bandwidth Powered Resonance Tube Actuators with Feedback Control,” Journal of
Sound and Vibration, Vol. 269, pp. 1031–1062.
Rampunggoon, P., and Mittal, R. (2002), “On the Virtual Aeroshaping of Synthetic Jets,”
Physics of Fluids, Vol. 14, No. 4, pp. 1533–1536.
Rao, N. M., Febg, J., Burdisso, R. A., and Ng, W. F. (2001), “Experimental Demonstration
of Active Flow Control to Reduce Unsteady Stator–Rotor Interaction,” AIAA Journal,
Vol. 39, No. 3, pp. 458–464.
REFERENCES 493

Rapoport, D., Fono, I., Cohen, K., and Seifert, A. (2002), “Closed-Loop Vectoring Control
of a Turbulent Jet Using Periodic Excitation,” AIAA Paper 2002-3073.
Rapoport, D., Fono, I., Cohen, K., and Seifert, A. (2003), “Closed-loop Vectoring Control
of a Turbulent Jet Using Periodic Excitation,” Journal of Propulsion and Power,
Vol. 19, No. 4, pp. 646–654.
Rasheed, A., Hornung, H. G., Fedorov, A. V., and Malmuth, N. D. (2002), “Experiments on
Passive Hypervelocity Boundary Layer Control Using a Ultrasonically Absorptive
Surface,” AIAA Journal, Vol. 40, No. 3, pp. 481–489.
Raspet, A. (1951), “Mechanisms of Automatic Trailing Edge Suction,” Mississippi State
College, Research Report 1, Engineering Research Station, 31 Dec. 1951.
Raspet, A. (1952), “Boundary Layer Studies on a Sailplane,” Aeronautical Engineering
Review, Vol. 11, No. 6, pp. 52–60.
Raspet, A., Cornish, J. J., III, and Bryant, G. D. (1956), “Delay of the Stall by Suction
through Distributed Perforations,” Aeronautical Engineering Review, Vol. 15, No. 8,
pp. 32–39.
Rathnasingham, R., and Breuer, K. S. (2003), “Active Control of Turbulent Boundary
Layers,” Journal of Fluid Mechanics, Vol. 495, pp. 209–233
Raverdy, B., Mary, I., and Sagaut, P. (2003), “High-Resolution Large-Eddy Simulation of
Flow around Low-Pressure Turbine Blade,” AIAA Journal, Vol. 41, No. 3, pp. 390–397.
Ravi, B. R., Mittal, R., and Najjar, F. M. (2004), “Study of Three-Dimensional Synthetic
Jet Flowfields Using Direct Numerical Simulation,” AIAA Paper 2004-0091.
Ravindran, S. S. (2000), “Reduced-Order Adaptive Controllers for Fluid Flows Using
POD,” Journal of Scientific Computing, Vol. 15, No. 4, pp. 457–478.
Rayleigh, L. (1945), The Theory of Sound, Dover Publications, New York.
Reau, N., and Tumin, A. (2002), “On Harmonic Perturbations in a Turbulent Mixing Layer,”
European Journal of Mechanics, B/Fluids, Vol. 21, No. 2, pp. 143–155.
Reddy, S. C., Schmid, P. J., and Henningson, D. S. (1993), “Pseudospectra of the
Orr–Sommerfeld Operator,” SIAM Journal of Applied Mathematics, Vol. 53, No. 1,
pp. 15–47.
Rehman, A., and Kontis, K. (2006), “Synthetic Jet Control Effectiveness on Stationary and
Pitching Airfoils,” Journal of Aircraft, Vol. 43, No. 6, pp. 1782–1798.
Reimann, D., Bloxham, M., Crapo, K. L., Pluim, J. D., and Bons, J. P. (2006), “Influence
of Vortex Generator Jet-Induced Transition on Separating Low-Pressure Turbine
Boundary Layers,” AIAA Paper 2006-2852.
Reimann, D., Pluim, J., and Bons, J. (2007), “Influence of Two-Dimensional vs. Discrete
Disturbances on Separating Low-Pressure Turbine Boundary Layers,” AIAA Paper
2007-525.
Reisenthel, P. (1988), Hybrid Instability in an Axisymmetric Jet with Enhanced Feedback,
Ph.D. Dissertation, Illinois Inst. Technology, Chicago, IL.
Rey, G., Banaszuk, A., and Gysling, D. (2003), “Active Control of Flutter in Turbomachinery
Using off Blade Actuators and Sensors: Experimental Results,” AIAA Paper 2003-
1258.
Reynolds, O. (1883), “An Experimental Investigation of the Circumstances which
Determine Whether the Motion of Water Shall be Direct or Sinuous, and of the Law of
Resistance in Parallel Channels,” Philosophical Transactions of the Royal Society,
Vol. 174, pp. 935–982.
Reynolds, T. S. (1983), Stronger Than a Hundred Men: A History of the Vertical Water
Wheel, Johns Hopkins Univ. Press, Baltimore, MA, pp. 14–17.
494 REFERENCES

Reynolds, O. (1894), “On the Dynamical Theory of Turbulent Incompressible Viscous


Fluids and the Determination of the Criterion,” Philosophical Transactions of the Royal
Society, A, Vol. 186, pp. 123–161.
Reynolds, W. C., and Carr, L. W. (1985), “Review of Unsteady, Driven, Separated Flows,”
AIAA Paper 1985-0527.
Reynolds, W. C., and Hussain, A. K .M. F. (1972), “The Mechanics of an Organized Wave
in Turbulent Shear Flow. Part 3. Theoretical Models and Comparisons with
Experiments,” Journal of Fluid Mechanics, Vol. 54, Part 2, pp. 263–288.
Reynolds, W. C., Parekh, D. E., Juvet, P. J. D., and Lee, M. J. D. (2003), “Bifurcating and
Blooming Jets,” Annual Review of Fluid Mechanics, Vol. 35, pp. 295–315.
Rhee, I., and Speyer, J. (1989), “A Game Theoretic Controller and its Relationship to
H• and Linear-Exponential-Gaussian Synthesis,” Proceedings of the 28th Conference
on Decision and Control, Vol. 2, pp. 909–915.
Rice, E. J., and Zaman, K. B. M. Q. (1987), “Control of Shear Flows by Artificial
Excitation,” NASA TM-100201 and AIAA Paper 87-2722.
Richards, G. A., and Straub, D. L. (2005), “Passive Control of Combustion Instabilities in
Stationary Gas Turbines,” Combustion Instabilities in Gas Turbine Engines, edited by
T. Lieuwen, and V. Yang, Vol. 210, AIAA Progress in Astronautics and Aeronautics,
pp. 533–580.
Richards, G. A., Janus, M., and Robey, E. H. (1999), “Control of Flame Oscillations with
Equivalence Ratio Modulation,” Journal of Propulsion and Power, Vol. 15, No. 2,
pp. 232–240.
Richards, G. A., McMillian, M. M., Gemmen, R. S., Rogers, W. A., and Cully, S. R. (2001),
“Issues for Low-Emission, Fuel-Flexible Power Systems,” Progress in Energy and
Combustion Science, Vol. 27, No. 2, pp. 141–169.
Richards, G., Rogbey, E., and Straub, D. (2003), “Control of Combustion Dynamics Using
Fuel System Impedance,” ASME GT-2003-3852.
Richards, G. A., Thornton, J. D., Robey, E. H., and Arellano, L. (2007), “Open-Loop Active
Control of Combustion Dynamics on a Gas Turbine Engine,” Journal of Engineering
for Gas Turbine and Power, Vol. 129, pp. 38–48.
Riley, A. J., Park, S., Dowling, A. P., Evesque, S., and Annaswamy, A. M. (2004), “Advanced
Closed-Loop Control on an Atmospheric Gaseous Lean-Premixed Combustor,” Journal
of Engineering for Gas Turbines and Power, Vol. 126, pp. 708–716.
Ritchie, B. D., Mujumdar, D. R., and Seitzman, J. M. (2000), “Mixing in Coaxial Jets
Using Synthetic Jet Actuators,” AIAA Paper 2000-0404.
Rivir, R., White, A., Carter, C., and Ganguly, B. (2004), “AC and Pulsed Plasma Flow
Control,” AIAA Paper 2004-0847.
Rizzetta, D. P., and Visbal, M. R. (2003a), “Numerical Investigations of Transitional Flow
through a Low-Pressure Turbine Cascade,” AIAA Paper 2003-3587.
Rizzetta, D. P., and Visbal, M. R. (2003b), “Large-Eddy Simulation of Supersonic Cavity
Flowfields Including Flow Control,” AIAA Journal, Vol. 41, No. 8, pp. 1452–1462.
Rizzetta, D. P., and Visbal, M. R. (2004), “Numerical Simulation of Separation Control for
a Transitional Highly-Loaded Low-Pressure Turbine,” AIAA Paper 2004-2204.
Rizzetta, D. P., and Visbal, M. R. (2005), “Numerical Simulation of Separation Control for
Transitional Highly Loaded Low-Pressure Turbines,” AIAA Journal, Vol. 43, No. 9,
pp. 1958–1967.
Rizzetta, D. P., and Visbal, M. R. (2006), “Direct Numerical Simulations of Flow Past an
Array of Distributed Roughness Elements,” AIAA Paper 2006-3527.
REFERENCES 495

Rizzetta, D. P., and Visbal, M. R. (2007), “Numerical Investigation of Plasma-Based Flow


Control for a Transitional Highly-Loaded Low-Pressure Turbine,” AIAA Paper
2007-0938.
Rizzetta, D. P., Visbal, M. R., and Stanek, M. J. (1999), “Numerical Investigation of
Synthetic-Jet Flowfields,” AIAA Journal, Vol. 37, No. 8, pp. 919–927.
Roberts, F. A. (1985), Effects of a Periodic Disturbance on Structure and Mixing in
Turbulent Shear Layers and Wakes. Ph.D. Thesis, California Inst. Techology, Pasadena,
CA.
Rockwell, D., and Naudascher, E. (1978), “Review-Self-Sustaining Oscillations of Flow
Past Cavities,” Journal of Fluids Engineering, Vol. 100, pp. 152–165.
Roos, F. W. (2000), “Microblowing: an Effective, Efficient Method of Vortex-Asymmetry
Management,” AIAA Paper 2000-4416.
Roshko, A. (1954a), “A New Hodograph for Free-Streamline Theory,” NACA TN-3168.
Roshko, A. (1954b), “On the Drag and Shedding Frequency of Two-Dimensional Bluff
Bodies,” NACA TM-3169.
Rossi, M. (1988), Acoustics and Electroacoustics, Artech House, Norwood, MA.
Rossiter, J. E. (1964), “Wind-Tunnel Experiments on the Flow over Rectangular Cavities
at Subsonic and Transonic Speeds,” Aeronautical Research Council R. & M.,
No. 3438.
Rossow V. J. (1999), “Lift-Generated Vortex Wakes of Subsonic Transport Aircraft,”
Progress in Aerospace Sciences, Vol. 35, No. 6, pp. 507–660.
Roth, J. D. (2003), Numerical Study of Active Flow Control Using Synthetic Jets, M.S.
Thesis, Iowa State University, Ames, IA.
Roth, J. R., Sherman, D. M., and Wilkinson, S. P. (1998), “Boundary Layer Control
with a One Atmosphere Uniform Glow Discharge Surface Plasma,” AIAA Paper
1998-328.
Roth, J. R., Sherman, D. M., and Wilkinson, S. P. (2000), “Electrohydrodynamic Flow
Control with a Glow-Discharge Surface Plasma,” AIAA Journal, Vol. 38, No. 7,
pp. 1166–1172.
Rouser, K. P. (2002), “Use of Dimples to Suppress Boundary Layer Separation on a Low
Pressure Turbine Blade,” AFIT Thesis AFIT/GAE/ENY/02-13.
Roussopoulos, K. (1993), “Feedback Control of Vortex Shedding at Low Reynolds
Numbers,” Journal of Fluid Mechanics, Vol. 248, pp. 267–296.
Roussopoulos, K., and Monkewitz, P. A. (1996), “Nonlinear Modeling of Vortex Shedding
Control in Cylinder Wakes,” Physica D, Vol. 97, pp. 264–273.
Rowley, C. W. (2005), “Model Reduction for Fluids Using Balanced Proper Orthogonal
Decomposition,” International Journal of Bifurcation Chaos, Vol. 15, No. 3,
pp. 997–1013.
Rowley, C. W., and Marsden, J. E. (2000), “Reconstruction Equations and the Karhunen–
Loève Expansion for Systems with Symmetry,” Physics D, Vol. 142, pp. 1–19.
Rowley, C. W., and Williams, D. R. (2006), “Dynamics and Control of High-Reynolds-
Number Flow over Open Cavities,” Annual Review of Fluid Mechanics, Vol. 38,
pp. 251–276.
Rowley, C. W., Colonius, T., and Basu, A. J. (2002), “On Self-Sustained Oscillations in
Two-Dimensional Compressible Flow over Rectangular Cavities,” Journal of Fluid
Mechanics, Vol. 455, pp. 315–346.
Rowley, C. W., Kevrekidis, I. G., Marsden, J. E., and Lust, K. (2003), “Reduction and Recon-
struction for Self-Similar Dynamical Systems,” Nonlinearity, Vol. 16, pp. 1257–1275.
496 REFERENCES

Rowley, C. W., Colonius, T., and Murray, R. M. (2004), “Model Reduction for Compressible
Flows Using POD and Galerkin Projection,” Physics D, Vol. 189, No. 1–2, pp. 115–129.
Rowley, C. W., Williams, D. R., Colonius, T., Murray, R. M., and MacMynowski, D. G.
(2006), “Linear Models for Control of Cavity Flow Oscillations,” Journal of Fluid
Mechanics, Vol. 546, pp. 317–330.
Roy, S. (2005), “Flow Actuation Using Radio Frequency in Partially-Ionized Collisional
Plasmas,” Applied Physics Letters, Vol. 86, No. 101502, pp. 1–3.
Roy, S., and Gaitonde, D. (2005), “Modeling Surface Discharge Effects of Atmospheric RF
on Gas Flow Control,” AIAA Paper 2005-0160.
Roy, S., Singh, K. P., and Gaitonde, D. (2006), “Dielectric Barrier Plasma Dynamics for
Active Control of Separated Flows,” Applied Physics Letters, Vol. 88, No. 12,
pp. 121501:1–3.
Rumsey, C. L. (2004), “Computation of a Synthetic Jet in a Turbulent Cross-Flow Boundary
Layer,” NASA TM-2004-213273.
Rumsey, C. L. (2007a), “Reynolds-Averaged Navier–Stokes Analysis of Zero Efflux Flow
Control over a Hump Model,” Journal of Aircraft, Vol. 44, No. 2, pp. 444–452.
Rumsey, C. L. (2007b), Proceedings of the 2004 Workshop on CFD Validation of Synthetic
Jets and Turbulent Separation Control, NASA CP-2007-214874.
Rumsey, C. L., and Gatski, T. B. (2001), “Recent Turbulence Model Advances Applied to
Multielement Airfoil Computations,” Journal of Aircraft, Vol. 38, No. 5, pp. 904–910.
Rumsey, C., Sanetrik, M., Biedron, R., Melson, N. D., and Parlette, E. (1996), “Efficiency
and Accuracy of Time-Accurate Turbulent Navier-Stokes Computations,” Computers
and Fluids, Vol. 25, No. 2, pp. 217–236.
Rumsey, C. L., Gatski, T. B., Sellers, W. L., Vatsa, V. N., and Viken, S. A. (2006), “Summary
of the 2004 Computational Fluid Dynamics Validation Workshop on Synthetic Jets,”
AIAA Journal, Vol. 44, No. 2, pp. 194–207.
Rumsey, C. L., Schaeffler, N. W., Milanovic, I. M., and Zaman, K. B. M. Q. (2007), “Time-
Accurate Computations of Isolated Circular Synthetic Jets in Crossflow,” Computers
and Fluids, Vol. 36, No. 6, pp. 1092–1105.
Rush, H. M. (1999), “Study of the Application of Separation Control by Unsteady Excitation
to Civil Transport Aircraft,” NASA CR-1999-209338.
Saad, Y. (1996), Iterative Methods for Sparse Linear Systems, PWS Publishing, Pacific
Grove, CA.
Saddoughi, S. (2004), “Synthetic Jet Actuators,” US Patent 6722581.
Sagaut, P. (2006), Large Eddy Simulation for Incompressible Flows, 3rd ed., Springer,
Berlin.
Sagaut, P., and Le, T. H. (1997), “Some Investigations on the Sensitivity of Large Eddy
Simulation,” Direct and Large-Eddy Simulation II, edited by J.-P. Chollet, P. R. Voke,
and L. Kleiser, Kluwer Academic Publishers, Dordrecht, pp. 81–92.
Saiyed, N. H., Mikkelsen, K. L., and Bridges, J. E. (2000), “Acoustics and Thrust of
Separate-Flow Exhaust Nozzles with Mixing Devices for High-Bypass Ratio Engines,”
AIAA Paper 2000-1961.
Salikuddin, M., Brown, W. H., and Ahuja, K. K. (1987), “Noise from a Circulation Control
Wing with Upper Surface Blowing,” Journal of Aircraft, Vol. 24, No. 1, pp. 55–64.
Samimy, M., Adamovich, L., Webb, B., Kastner, J., Hileman, J., Keshav, S., and Palm, P.
(2004a), “Development and Characterization of Plasma Actuators for High-Speed Jet
Control,” Experiments in Fluids, Vol. 37, No. 4, pp. 577–588.
REFERENCES 497

Samimy, M., Adamovich, I., Webb, B., Kastner, J., Hileman, J., Keshav, S., and Palm, P.
(2004b), “Development and Application of Localized Arc Filament Plasma Actuators
for Jet Flow and Noise Control,” AIAA Paper 2004-0184.
Samimy, M., Adamovich, I., Kim, J.-H., Webb, B., Keshav, S., and Utkin, Y. (2004c),
“Active Control of High Speed Jets using Localized Arc Filament Plasma Actuators,”
AIAA Paper 2004-2130.
Samimy, M., Kim, J.-H., Kastner, J., Adamovich, I., and Utkin, Y. (2007a), “Active Control
of a Mach 0.9 Jet for Noise Mitigation Using Plasma Actuators,” AIAA Journal, Vol. 45,
No. 4, pp. 890–901.
Samimy, M., Kim, J.-H., Kastner, J., and Adamovich, I. (2007b), “Noise Mitigation in
High Speed and High Reynolds Number Jets Using Plasma Actuators,” AIAA Paper
2007-3622.
Sampatacos, E. P., Morger, K. M., and Logan, A. H. (1983), “NOTAR: The Viable
Alternative to a Tail Rotor,” AIAA Paper 1983-2527.
Sandham, N. D. (2001), “A Review of Progress on Direct and Large-Eddy Simulation,”
Modern Simulation Strategies for Turbulent Flow, edited by B. J. Geurts and
R. T. Edwards, Philadelphia, PA, pp. 1–20.
Saric, W. S. (1994), “Goertler Vortices,” Annual Review of Fluid Mechanics, Vol. 26,
pp. 379–409.
Saric, W. S., and Reed, H. L. (2002), “Supersonic Laminar Flow Control on Swept Wings
Using Distributed Roughness,” AIAA Paper 2002-0147.
Saric, W. S., and Reed, H. L. (2003), “Crossflow Instabilities—Theory and Technology,”
AIAA Paper 2003-0771.
Saric, W. S., Carrillo, R. B., and Reibert, M. S. (1998), “Leading-Edge Roughness as a
Transition Control Mechanism,” AIAA Paper 1998-0781.
Saric, S., Jakirlic, S., Djugum, A., and Tropea, C. (2006), “Computational Analysis of a
Locally Forced Flow over a Wall-Mounted Hump at High-Re Number,” International
Journal of Heat and Fluid Flow, Vol. 27, No. 4, pp. 707–720.
Sarpotdar, S., Raman, G., and Cain, A. B. (2005), “Powered Resonance Tubes: Resonance
Characteristics and Actuation Signal Directivity,” Experiments in Fluids, Vol. 39,
No. 6, pp. 1084–1095.
Sawyer, S., and Fleeter, S. (2000), “Active Control of Discrete-Frequency Turbomachinery
Noise Using a Rotary-Valve Actuator,” Journal of Engineering for Gas Turbines and
Power, Vol. 122, pp. 226–232.
Sawyer, S., Fleeter, S., and Simonich J. (1997), “Active Control of Discrete-Frequency
Noise Generated by Rotor–Stator Interactions,” AIAA Paper 97-1663.
Schadow, K. C., and Gutmark, E. (1992), “Combustion Instability Related to Vortex
Shedding in Dump Combustors and Their Passive Control,” Progress in Energy and
Combustion Science, Vol. 18, pp. 117–132.
Schaeffler, N. W., and Jenkins, L. N. (2006), “Isolated Synthetic Jet in Crossflow:
Experimental Protocols for a Validation Dataset,” AIAA Journal, Vol. 44, No. 12,
pp. 2846–2856.
Schaeffler, N. W., Hepner, T. E., Jones, G. S., and Kegerise, M. A. (2002), “Overview of
Active Flow Control Actuator Development at NASA Langley Research Center,”
AIAA Paper 2002-3159.
Scheeper, P. R., van der Donk, A. G. H., Olthuis, W., and Bergveld, P. (1994), “A Review
of Silicon Microphones,” Sensors and Actuators A, Vol. 44, pp. 1–11.
498 REFERENCES

Schewe, G. (1983), “On the Forced Fluctuations Acting on a Circular Cylinder in a Cross-
Flow from Sub-Critical to Trans-Critical Reynolds Number,” Journal of Fluid
Mechanics, Vol. 133, pp. 265–285.
Schilz, W. (1965), “Experimentelle Untersuchungen zur Akustischen Beeinussung der
Strömungsgrenzschicht in Luft,” Acustica, Vol. 16, pp. 208–223.
Schlichting, H. (1979), Boundary Layer Theory, McGraw-Hill, New York.
Schluter, J. U. (2001), “Large-Eddy Simulation of Combustion Instability Suppression by
Static Turbulence Control,” Center for Turbulence Research Annual Brief, pp. 119–127.
Schmid, P. (2000), “Linear Stability and Bypass Transition is Shear Flows,” Physics of
Plasmas, Vol. 7, No. 5, pp. 1788–1794.
Schmid, P. J., and Henningson, D.S. (1992), “A New Mechanism for Rapid Transition
Involving a Pair of Oblique Waves,” Physics of Fluids A, Vol. 4, No. 9, pp. 1986–1989.
Schmid, P. J., and Henningson, D. S. (2001), Stability and Transition in Shear Flows,
Springer, Berlin.
Schmidt, M. A., Howe R. T., Senturia S. D., and Haritonidis J. H. (1988), “Design and
Calibration of a Microfabricated Floating-Element Shear-Stress Sensor,” IEEE
Transactions on Electronic Devices, ED-35, pp. 750–757.
Schöber, M., Obermeier, E., Pirskawetz, S., and Fernholz, H.-H. (2004), “A MEMS Skin-
Friction Sensor for Time-Resolved Measurements in Separated Rows,” Experiments in
Fluids, Vol. 36, pp. 593–599.
Schrenk, O. (1928), “Tragfluegel mit Grenzschichtabsaugung,” Luftfahrtforschung 2, p. 49.
Schrenk, O. (1931), “Versuche an einim Absaugflugeln,” Z. Flugtech. Motorlufts. Vol. 22,
p. 259.
Schrenk, O. (1940), “Grenzschichtabsaugung,” Luftwissen, Vol. 7, p. 209.
Schubauer, G. B., and Klebanoff, P. S. (1955), “Contributions on the Mechanics of
Boundary-Layer Transition,” NACA Tech. Note 3489.
Schubauer, G. B., and Klebanoff, P. S. (1956), “Contributions to the Mechanics of
Boundary-Layer Transition,” NACA Tech. Rept. 1289.
Schubauer, G. B., and Skramstad, H. K. (1948), “Laminar Boundary Layer Oscillations
and Stability of Laminar Flow,” National Bureau of Standards Research Paper 1772
and NACA Rept. 909.
Schulten, J. B. H. M. (2001), “Active Control of Rotor–Stator Interaction Noise through
Vibrating Vanes,” AIAA Paper 2001-2151.
Schwendemann, M., and Sanders, B. (1982), “Tangential Blowing for Control of
Strong Normal Shock-Boundary-Layer Interactions on Inlet Ramps,” AIAA Paper
82-1082.
Seddon, J. J., and Goldsmith, E. L. (1999), Intake Aerodynamics, 2nd ed., AIAA Education
Series, New York.
Seeley, C. E., Arik, M., Hedeen, R., Utturkar, Y., Wetzel, T., and Shih, M.-Y. (2006),
“Coupled Acoustic and Heat Transfer Modeling of a Synthetic Jet,” AIAA Paper
2006-1879.
Seifert, A., and Pack, L. G. (1999), “Oscillatory Control of Separation at High Reynolds
Numbers”, AIAA Journal, Vol. 37, No. 9, pp. 1062–1071.
Seifert, A., and Pack, L. G. (2002), “Active Control of Separated Flow on a Wall-Mounted
Hump at High Reynolds Numbers,” AIAA Journal, Vol. 40, No. 7, pp. 1363–1372.
Seifert, A., and Pack, L. G. (2003a) “Effects of Compressibility and Excitation Slot
Location on Active Separation Control at High Reynolds Numbers,” Journal of Aircraft,
Vol. 40, No. 1, pp. 110–119.
REFERENCES 499

Seifert, A., and Pack, L. G. (2003b), “Effects of Sweep on Active Separation Control at
High Reynolds Numbers,” Journal of Aircraft, Vol. 40, No. 1, pp. 120–126.
Seifert, A., and Pack-Melton, L. (2006), “Identification and Control of Turbulent Boundary
Layer Separation”, 100 Years of Boundary Layer Research and L. Prandtl Centennial
Lecture, edited by G. E. A. Meier, K. R. Sreenivasan, and H. J. Heinemann, Springer-
Verlag, Berlin, pp. 199–208.
Seifert, A., Bachar, T., Koss, D., Shepshelovits, M., and Wygnanski, I. (1993), “Oscillatory
Blowing—a Tool to Delay Boundary-Layer Separation,” AIAA Journal, Vol. 31, No.
11, pp. 2052–2060.
Seifert, A., Darabi, A., and Wygnanski, I. (1996), “Delay of Airfoil Stall by Periodic
Excitation,” Journal of Aircraft, Vol. 33, No. 4, pp. 691–699.
Seifert, A., Eliahu, S., Greenblatt, D., and Wygnanski, I. (1998), “Use of Piezoelectric
Actuators for Airfoil Separation Control (TN),” AIAA Journal, Vol. 36, No. 8,
pp. 1535–1537.
Seifert, A., Bachar, T., Wygnanski, I., Kariv, A., Cohen, H., and Yoeli, R. (1999), “Application
of Active Separation Control to a Small UAV,” Journal of Aircraft, Vol. 36, No. 2,
pp. 474–477.
Seifert, A., Greenblatt, D., and Wygnanski, I. (2004), “Active Separation Control: A Review
of Reynolds and Mach Numbers Effects,” Aerospace Science and Technology, Vol. 8,
pp. 569–582.
Seiner, J. M. (1998), “A New Rational Approach to Jet Noise Reduction,” Theoretical and
Computational Fluid Dynamics, Vol. 10, No. 1–4, pp. 373–383.
Seiner, J. M., Ukeiley, L., and Jansen, B. (2005), “Aero-Performance Efficient Noise
Reduction for the F404-400 Engine,” AIAA Paper-2005-3048.
Selby, G. V., Lin, J. C., and Howard, F. G. (1992), “Control of Low-Speed Turbulent
Separated Flow Using Jet Vortex Generators,” Experiments in Fluids, Vol. 12,
pp. 394–400.
Sell, J. (1997), Cascade Testing to Assess the Effectiveness of Mass Addition/Removal Wake
Management Strategies for Reduction of Stator Rotor Interaction Noise, M.S. Thesis,
Massachusetts Inst. Technology, Cambridge, MA.
Selle, L., Benoit, L., Poinsot, T., Nicoud, F., and Krebs, W. (2006), “Joint Use of
Compressible Large-Eddy Simulation and Helmholtz Solvers for the Analysis of
Rotating Modes in an Industrial Swirled Burner,” Combustion and Flame, Vol. 145,
No. 1–2, pp. 194–205.
Sellers, W. L., III, Jones, G. S., and Moore, M. D. (2002), “Flow Control Research at
NASA Langley in Support of High Lift Augmentation,” AIAA Paper 2002-6006.
Sharma, O. (1998), “Impact of Reynolds Number on LP Turbine Performance,” NASA
CP-1998-206958, Minnowbrook II Workshop on Boundary Layer Transition in
Turbomachines, edited by J. E. LaGraff and D. E. Ashpis, pp. 65–69.
Sharma, R. N. (2007), “Fluid-Dynamics-Based Analytical Model for Synthetic Jet
Actuation,” AIAA Journal, Vol. 45, No. 8, pp. 1841–1847.
Sharma, O. P., Ni, R. H., and Tanrikut, S. (1994), “Unsteady Flow in Turbines—Impact on
Design Procedure,” Turbomachinery, AGARD Lecture Series 195.
Sharma, A. S., Abdessemed, N., Sherwin, S. J., and Theofilis, V. (2008), “Optimal Growth
of Linear Perturbations in Low Pressure Turbine Flows,” IUTAM Symposium on Flow
Control and MEMS, edited by J. F. Morrison, D. M. Birch and P. Lavoie, pp. 339–343
Shaw, L., Smith, B. R., and Saddoughi, S. (2005), “Active Control of a Pod Wake-Mid-
Scale Application,” AIAA Paper 2005-799.
500 REFERENCES

Shaw, L., Smith, B. R., and Saddoughi, S. (2006), “Full-Scale Flight Demonstration of
Active Control of a Pod Wake,” AIAA Paper 2006-3185.
Sheplak, M., Cattafesta, L., Nishida, T., and McGinley, C. B. (2004), “MEMS Shear Stress
Sensors: Promise and Progress,” AIAA Paper 2004-2606.
Sheplak, M., Cattafesta, L., and Tian, Y. (2008), “Micromachined Shear Stress Sensors for
Flow Control Applications,” IUTAM Symposium on Flow Control with MEMS, Royal
Geographical Society London, September 2006. IUTAM Bookseries, Vol. 7, Morrison, J. F.,
Birch, D. M., Lavoie, P. (Eds), pp. 67–76.
Shepshelovich, M., Koss, D., Wygnanski, I., and Seifert, A. (1989), “An Experimental
Evaluation of a Low-Reynolds Number High-Lift Airfoil with Vanishingly Small
Pitching Moment,” AIAA Paper 1989-538.
Shin, J., Narayanaswamy V., Rajay, L. L., and Clemensz, N. (2006), “Generation of Plasma
Induced Flow Actuation by DC Glow-like Discharge in a Supersonic Flow,” AIAA
Paper 2006-169.
Shinjo, J., Masuyama, S., Mizobuchi, Y., and Ogawa, S. (2007), “Study on Flame Dynamics
with Secondary Fuel Injection Control by Large Eddy Simulation,” Combustion and
Flame, Vol. 150, pp. 277–291.
Shyy, W., Jayaraman, B., and Andersson, A. (2002), “Modeling of Glow-Discharge Induced
Flow Dynamics,” Journal of Applied Physics, Vol. 92, No. 11, pp. 6434–6443.
Siegel, S. G., McLaughlin, T. E., and Morrow, J. A. (2001), “PIV Measurements on a Delta
Wing with Periodic Blowing and Suction,” AIAA Paper 2001-2436.
Siegel, S., Cohen, K., and McLaughlin, T. (2003), “Feedback Control of a Circular Cylinder
Wake in Experiment and Simulation,” AIAA Paper 2003-3571.
Siegel, S. G., Cohen, K., and McLaughlin, T. (2006), “Numerical Simulations of a Feedback
Controlled Circular Cylinder Wake,” AIAA Journal, Vol. 44, No. 6, pp. 1266–1276.
Sigurdson, L. W. (1995), “The Structure and Control of a Turbulent Reattaching Flow,”
Journal of Fluid Mechanics, Vol. 298, pp. 139–165.
Sigurdson, L. W., and Roshko, A. (1985), “Controlled Unsteady Excitation of a Reattaching
Flow,” AIAA Paper 1985-552.
Simonich, J., Lavrich, P., Sofrin, T., and Topol, D. (1993), “Active Aerodynamic Control of
Wake-Airfoil Interaction Noise—Experiment,” AIAA Journal, Vol. 31 No. 10,
pp. 1761–1768.
Singh, K. P., and Roy, S. (2007), “Modeling Plasma Actuators with Air Chemistry for
Effective Flow Control,” Journal of Applied Physics, Vol. 101, p. 123308.
Sirovich, L. (1987), “Turbulence and the Dynamics of Coherent Structures, Parts I–III,”
Quarterly Applied Mathematics, Vol. XLV, No. 3, pp. 561–590.
Skogestad, S., and Postlethwaite, I. (1996), Multivariable Feedback Control, John Wiley,
Chichester, UK.
Slomski, J. F., Gorski, J. J., Miller, R. W., and Marino, T. A. (2002), “Numerical Simulation
of Circulation Control Airfoils as Affected by Different Turbulence Models,” AIAA
Paper 2002-0851.
Slomski, J. F., Chang, P. A., and Arunajatesan, S. (2006), “Large Eddy Simulation of a
Circulation Control Airfoil,” AIAA Paper 2006-3011.
Smith, C. S. (1954), “Piezoresistance Effect in Germanium and Silicon,” Physics Reviews,
Vol. 94, No. 1, pp. 42–49.
Smith, B. R. (2000), “Computational Simulation of Active Control of Cavity Acoustics,”
AIAA Paper 2000-1927.
REFERENCES 501

Smith, B. (2007), “Unit 4 Lecture Notes—Modern Active Flow Control,” AIAA Short
Course, Reno.
Smith, B. L., and Glezer, A. (1994), “Vectoring of a High Aspect Ratio Air Jet using Zero-
Net-Mass-Flux Control Jet,.” Bulletin of the American Physical Society, Vol. 39,
p. 1894.
Smith, B. L., and Glezer, A. (1998), “The Formation and Evolution of Synthetic Jets,”
Physics of Fluids, Vol. 10, No. 9, pp. 2281–2297.
Smith B. L., and Glezer, A. (2002), “Jet Vectoring Using Synthetic Jets,” Journal of Fluid
Mechanics, Vol. 458, pp. 1–34.
Smith, A. N., Babinsky, H., Dhanasekaran, Savill, A. M., and Dawes, W. N. (2003),
“Computational Investigation of Groove Controlled Shock Wave/Boundary Layer
Interaction,” AIAA Paper 2003-0446.
Smith, A. N., Babinsky, H., Fulker, J. L., and Ashill, P. R. (2004), “Shock-Wave/Boundary
Layer Interaction Control using Streamwise Slots in Transonic Flows,” Journal of
Aircraft, Vol. 41, No. 3, pp. 540–546.
Smith, D. M., Dickey, E., and VonKlein, T. (2006), “The ADVINT Program,” AIAA Paper
2006-2854.
Sondergaard, R., Bons, J. P., and Rivir, R. B. (2002a), “Control of Low-Pressure Turbine
Separation Using Vortex Generator Jets,” Journal of Propulsion and Power, Vol. 18,
No. 4, pp. 889–895.
Sondergaard, R., Bons, J. P., Sucher, M., and Rivir, R. B. (2002b), “Reducing Low-Pressure
Turbine Stage Blade Count using Vortex Generator Jet Separation Control,” Proceedings
of ASME Turbo Expo 2002, Amsterdam, June.
Sondergaard, R., Rivir, R., Bons, J., and Yurchenko, N. (2004), “Control of Separation in
Turbine Boundary Layers,” AIAA Paper 2004-2201.
Sosa, R., Artana, G., Moreau, E., and Touchard, G. (2006), “Flow Control with EHD
Actuators in Middle Post Stall Regime,” Journal of the Brazilian Society of Mechanics,
Science and Engineering, Vol. XXVIII, No. 2, pp. 200–207.
Spalart, P. R. (1998), “Airplane Trailing Vortices,” Annual Review of Fluid Mechanics,
Vol. 30, pp. 107–138.
Spalart, P. R., and Allmaras, S. R. (1994), “A One-Equation Turbulence Model for
Aerodynamic Flows,” La Recherche Aerospatiale, Vol. 1, pp. 5–21.
Spalart, P. R., and Shur, M. (1997), “On the Sensitization of Turbulence Models to Rotation
and Curvature,” Aerospace Science and Technology, Vol. 1, No. 5, pp. 297–302.
Spalart, P. R., Jou, W.-H., Strelets, M., and Allmaras, S. R. (1997), “Comments on the
Feasibility of LES for Wings, and on a Hybrid RANS/LES Approach,” Advances in
DNS/LES, edited by C. Liu, and Z. Liu, Greyden Press, Columbus, OH, pp. 137–148.
Spalart, P., Hedges, L., Shur, M., and Travin, A. (2003), “Simulation of Active Flow Control
on a Stalled Airfoil,” Flow, Turbulence and Combustion, Vol. 71, pp. 361–373.
Spalart, P. R., Deck, S., Shur, M. L., Squires, K. D., Strelets, M. K., and Travin, A. (2006),
“A New Version of Detached-Eddy Simulation, Resistant to Ambiguous Grid Densities,”
Theoretical and Computational Fluid Dynamics, Vol. 20, pp. 181–195.
Speziale, C. G. (1998), “Turbulence Modeling for Time-Dependent RANS and VLES: A
Review,” AIAA Journal, Vol. 36, No. 2, pp. 173–184.
Sreenivasan, K., and White, C. M. (2000), “The Onset of Drag Reduction by Dilute Polymer
Additives, and the Maximum Drag Reduction Asymptote,” Journal of Fluid Mechanics,
Vol. 409, pp. 149–164.
502 REFERENCES

Stack, J. (1933), “The NACE High-Speed Wind Tunnel Test of Six Propeller Sections,”
NACA TR-463.
Stalker, A. (2004), “The Effects of Krueger Flaps and Periodic Perturbations on the
Download Alleviation of a Typical Tiltrotor Aircraft,” M.Sc. Thesis, Department of
Aerospace and Mechanical Engineering, The Univ. Arizona Tucson, AZ.
Stalnov, O., Palei, V., Fono, I., Cohen, K., and Seifert, A. (2007), “Experimental Estimation
of a ‘D’ Shaped Cylinder Wake using Body Mounted Sensors,” Experiments in Fluids,
Vol. 44, No. 4, pp. 531–542.
Stanek, M., Raman, G., Kibens, V., Ross, J., Peto. J., Odedra, J. (2001), “Suppression of
Cavity Resonance using High Frequency Forcing—the Characteristic Signature of
Effective Devices,” AIAA Paper 2001-2128.
Stanek, M. J., Raman, G., Ross, J. A., Odedra, J., Peto, J., Alvi, F., and Kibens, V. (2002a),
“High Frequency Acoustic Suppression—the Role of Mass Flow, the Notion of
Superposition, and the Role of Inviscid Instability—a New Model (Part II),” AIAA
Paper 2002-2404.
Stanek, M. J., Sinha, R., Seiner, J., Pierce, B., and Jones, M. (2002b), “High Frequency
Flow Control—Suppression of Aero-Optics in Tactical Directed Energy Beam
Propagation and the Birth of a New Model (Part I),” AIAA Paper 2002-2272.
Stanek, M. J., Raman, G., Ross, J. A., Odedra, J., Peto, J., Alvi, F. S., and Kibens, V. (2003),
“High Frequency Acoustic Suppression—the Mystery of the Rod-in-Crossflow
Revealed,” AIAA Paper 2003-0007.
Stanewsky, E. (2001), “Adaptive Wing and Flow Control Technology,” Progress in
Aerospace Sciences, Vol. 37, pp. 583–667.
Steele, R., Cowell, L. H., Cannon, S., and Smith, C. (2000), “Passive Control of Combustion
Instability in Lean Premixed Combustors,” Journal of Engineering for Gas Turbines
and Power, Vol. 122, pp. 412–419.
Stengel, R. F. (1994), Optimal Control and Estimation, Dover Publications, New York.
Stieger, R. D., and Hodson, H. P. (2003), “The Transition Mechanism of Highly-Loaded LP
Turbine Blades,” ASME Turbo Expo 2003: Power for Land, Sea, and Air, GT2003-
38304.
Stone, C., and Menon, S. (2002), “Adaptive Swirl Control of Combustion Dynamics in
Gas Turbine Combustors,” Proceedings of the Combustion Institute, Vol. 29, pp.
155–160.
Stone, C., and Menon, S. (2003), “Open Loop Control of Combustion Instabilities in a
Model Gas Turbine Combustor,” Journal of Turbulence, Vol. 4, No. 4, pp. 1–32.
Stratford, B. S. (1956) “An Experimental Flow with Zero Skin Friction Throughout its
Region of Pressure Rise,” Journal of Fluid Mechanics, Vol. 5, pp. 17–35.
Straub, F. K., and Hassan, A. A. (1996), “Aeromechanic Considerations in the Design of a
Rotor with Smart Material Actuated Trailing Edge Flaps,” 52nd Annual Forum of the
American Helicopter Society, Washington, DC.
Straub, D. L., Casleton, K., Lewis, R., Sidwell, T., Maloney, D. J., and Richards, G. A.
(2005), “Assessment of Rich-Burn, Quick-Mix, Lean-Burn Trapped Vortex Combustor
for Stationary Gas Turbine,” Journal of Engineering for Gas Turbines and Power,
Vol. 127, pp. 36–41.
Strykowski, P. J., and Krothapalli, A. (1993), “The Countercurrent Mixing Layer: Strategies
for Shear-Layer Control,” AIAA Paper 93-3260.
Strykowski, P. J., and Sreenivasan, K. R. (1990), “On the Formation and Suppression of
Vortex ‘Shedding’ at Low Reynolds Numbers,” Journal of Fluid Mechanics, Vol. 218,
pp. 71–107.
REFERENCES 503

Strykowski, P. J., Krothapalli, A., and Wishart, D. (1993), “Enhancement of Mixing in


High-Speed Heated Jets Using a Counterflow Nozzle,” AIAA Journal, Vol. 31, No. 11,
pp. 2033–2038.
Strykowski, P. J., Krothapalli, A., and Forliti, D. J. (1996), “Counterflow Thrust Vectoring,”
AIAA Journal, Vol. 34, No. 11, pp. 2306–2314.
Strykowski, P. J., Schmidt, G. F., Alvi, F. S., and Krothapalli, A. (1997), “Vectoring Thrust
Using Confined Countercurrent Shear Layers,” AIAA Paper 97-1997.
Sumitani, Y., and Kasagi, N. (1995), “Direct Numerical Simulation of Turbulent
Transport with Uniform Wall Injection and Suction,” AIAA Journal, Vol. 33, No. 7,
pp. 1220–1228.
Suponitsky, V., Avital, E., and Gaster, M. (2005), “On Three-Dimensionality and Control
of Incompressible Cavity Flows,” Physics of Fluids, Vol. 17, 104103.
Surana, A., Grunberg, O., and Haller, G. (2006), “Exact Theory of Three-Dimensional Flow
Separation. Part I. Steady Separation,” Journal of Fluid Mechanics, Vol. 564, pp. 57–103
Sureshkumar, R., and Beris, A. N. (1995), “Effect of Artificial Stess Diffusivity on the
Stability of Numerical Calculations and the Flow Dynamics of Time-Dependent
Viscoelastic Flows,” Journal of Non-Newtonian Fluid Mechanics, Vol. 60, pp. 53–80.
Sureshkumar, R., Beris, A. N., and Handler, R. A. (1997), “Direct Numerical Simulation of
the Turbulent Channel Flow of a Polymer Solution,” Physics of Fluids, Vol. 9,
pp. 743–755.
Sutliff, D. L. (2005), “Broadband Noise Reduction of a Low-Speed Fan Noise Using
Trailing Edge Blowing,” AIAA Paper 2005-3030.
Sutliff, D. L., Tweedt, D. L., Fite, E. B., and Envia, E. (2002), “Low-Speed Fan Noise
Reduction with Trailing-Edge Blowing,” AIAA Paper 2002-2492.
Suzen, Y. B., and Huang, P. G. (2005), “Numerical Simulation of Unsteady Wake-Blade
Interactions in Low-Pressure Turbine Flows Using an Intermittency Transport
Equation,” Journal of Turbomachinery, Vol. 127, No. 3, pp. 431–444.
Suzen, Y. B., Huang, P. G., and Jacob, J. D. (2005), “Numerical Simulations of Plasma
Based Flow Control Applications,” AIAA Paper 2005-4633.
Suzuki, T., and Colonius, T. (2006), “Instability Waves in a Subsonic Round Jet Detected
Using a Near-Field Phased Microphone Array,” Journal of Fluid Mechanics, Vol. 565,
pp. 197–226.
Suzuki, H., Kasagi, N., Suzuki, Y., and Shima, H. (1999), “Manipulation of a Round Jet
with Electromagnetic Flap Actuators,” IEEE MEMS ’99, pp. 534–540.
Suzuki, T., Colonius, T., and Pirozzoli, S. (2004), “Vortex Shedding in a Two-Dimensional
Diffuser: Theory and Simulation of Separation Control by Periodic Mass Injection,”
Journal of Fluid Mechanics, Vol. 520, pp. 187–213.
Swanson, R. C., and Rumsey, C. L. (2006), “Numerical Issues for Circulation Control
Calculations,” AIAA Paper 2006-3008.
Swanson, R. C., Rumsey, C. L., and Anders, S. G. (2005), “Progress Towards Computational
Method for Circulation Control Airfoils,” AIAA Paper 2005-0089.
Swanson, R. C., Rumsey, C. L., and Anders, S. G. (2006), “Aspects of Numerical Simulation
of Circulation Control Airfoils,” Applications of Circulation Control Technology,
edited by R. D. Joslin, and G. S. Jones, Vol. 214, AIAA Progress in Aeronautics and
Astronautics Series, Reston, VA, pp. 469–498.
Tabor, M., and de Gennes, P. G. (1986), “A Cascade Theory of Drag Reduction,” Europhysics
Letters, Vol. 2, pp. 519–522.
Tacina, R., Mao, C. P., and Wey, C. (2003), “Experimental Investigation of a Multiplex
Fuel Injector Module for Low Emission Combustors,” AIAA Paper 2003-0827.
504 REFERENCES

Tam, C. K. W., and Morris, P. J. (1980), “The Radiation of Sound by the Instablity Waves
of a Compressible Plane Turbulent Shear Layer,” Journal of Fluid Mechanics, Vol. 98,
pp. 349–381.
Tam, C. K., Golebiowski, M., and Seiner, J. M. (1996), “On the Two Components of
Turbulent Mixing Noise from Supersonic Jets,” AIAA Paper 1996-1716.
Taneda, S. (1978), “Visual Observations of the Flow Past a Circular Cylinder Performing a
Rotary Oscillation,” Journal of the Physical Society of Japan, Vol. 45, p. 1038.
Tang, H., Zhong, S., Jabbal, M., Garcillan, L., Guo, F., Wood, N., and Warsop, C. (2007),
“Towards the Design of Synthetic-Jet Actuators for Full-Scale Flight Conditions—
Part 2: Low-Dimensional Performance Prediction Models and Actuator Design
Method,” Flow, Turbulence and Combustion, Vol. 78, No. 3, pp. 309–329.
Tani, I. (1977), “History of Boundary Layer Theory,” Annual Review of Fluid Mechanics,
Vol. 9, pp. 87–111.
Tanna, H. K. (1977), “An Experimental Study of Jet Noise, Part II: Shock Associated
Noise,” Journal of Sound and Vibration, Vol. 50, pp. 429–444.
Tanna, H. K., and Ahuja, K. K. (1985), “Tone Excited Jets, Part 1: Introduction,” Journal
of Sound and Vibration, Vol. 102, No. 1, pp. 57–61.
Taubert, L., Kjellgren, P., and Wygnanski, I. (2002), “Generic Bluff Bodies with
Undetermined Separation Location,” AIAA Paper 2002-3068.
Tennekes, H., and Lumley, J. L. (1972), A First Course in Turbulence, MIT Press,
Cambridge, MA.
Theofilis, V. (2000), “Global Linear Instability in Laminar Separated Boundary Layer
Flow,” IUTAM Laminar-Turbulent Symposium V, edited by W. Saric and H. Fasel,
Sedona, AZ, pp. 663–668.
Theofilis, V. (2003), “Advances in Global Linear Instability Analysis of Nonparallel and
Three-Dimensional Flows,” Progress in Aerospace Sciences, Vol. 39, pp. 249–315.
Theofilis, V., and Sherwin, S. J. (2004), “Global Instability and Control of Low-Pressure
Turbine Flows,” NASA TM-2004-212913, edited by J. E. LaGraff and D. E. Ashpis,
pp. 31–32.
Theofilis, V., Hein, S., and Dallmann, U. (2000), “On the Origins of Unsteadiness and
Three-Dimensionality in a Laminar Separation Bubble,” Philosophical Transactions of
the Royal Society of London A, Vol. 358, pp. 3229–3246.
Thiele, U., Vega, J. M., and Knobloch, E. (2006), “Long-Wave Marangoni Instability with
Vibration,” Journal of Fluid Mechanics, Vol. 546, pp. 61–87.
Thiruchengode, M., Nair, S., Olsen, R., Neumeier, Y., Meyers, A., Seitzman, J., Jagoda, J.,
Lieuwen, T., and Zinn, B. T. (2004), “Blowout Control in Turbine Engine Combustors,”
AIAA Paper 2004-0637.
Thomas, A. (1983), “The Control of Boundary-Layer Transition Using a Wave-Superposition
Principle,” Journal of Fluid Mechanics, Vol. 137, pp. 233–250.
Thomas, R. H., Burdisso, R. A., Fuller, C. R., and O’Brien, W. F. (1993), “Preliminary
Experiments on Active Control of Fan Noise from a JT15D Turbofan Engine,” Journal
of Sound and Vibration, Vol. 161, No. 3, pp. 532–537.
Tian, Y., Cattafesta, L., and Mittal, R. (2006a), “Adaptive Control of Separated Flow,”
AIAA Paper 2006-1401.
Tian, Y., Song, Q., and Cattafesta, L. (2006b), “Adaptive Feedback Control of Flow
Separation,” AIAA Paper 2006-3016.
Tietjens, O. (1922), “Beiträge zur Entstehung der Turbulenz,” Dissertation, Göttingen,
Germany.
REFERENCES 505

Timor, I., Ben-Hamou, E., Guy, Y., and Seifert, A. (2007), “Maneuvering Aspects and 3D
Effects of Active Airfoil Flow Control,” Flow, Turbulence and Combustion, Special
Issue on Air-Jet Actuators and their Use for Flow Control, Vol. 78, pp. 429–443.
Tindell R. H. (1987), “Highly Compact Inlet Diffuser Technology,”AIAA Paper 87-1747.
Tindell, R., and Willis, B. (1997), “Experimental Investigation of Blowing for Controlling
Oblique Shock/Boundary Layer Interactions,” AIAA Paper 97-2642.
Tinney, C., and Glauser, M. N. (2007), “The Modified Complementary Technique Applied
to the Mach 0.85 Axisymmetric Jet for Noise Prediction,” AIAA Paper 2007-3663.
Tinney, C. E., Glauser M. N., and Ukeiley, L. S. (2005), “The Evolution of the Most
Energetic Modes in a High Subsonic Mach number Turbulent Jet,” AIAA Paper 2005-
0417.
Tinney, C., Jordan, P., Guitton, A., Delville, J., and Coiffet, F. (2006), “A Study in the Near
Pressure Field of Co-Axial Subsonic Jets,” AIAA Paper 2006-2589.
Tokumaru P. T., and Dimotakis, P. E. (1991), “Rotary Oscillation Control of a Cylinder
Wake,” Journal of Fluid Mechanics, Vol. 224, pp. 77–90.
Tokumaru, P. T., and Dimotakis, P. E. (1993), “The Lift of a Cylinder Executing Rotary
Motions in a Uniform Flow,” Journal of Fluid Mechanics, Vol. 255, pp. 1–10.
Tollmien, W. (1929), “Über die Entstehung der Turbulenz, 1. Mitteilung,” Nachr. Acad.
Wiss., Göttingen, Math. Phys., pp. 21–44.
Tollmien, W., and Grohne, D. (1961), “The Nature of Transition,” Boundary Layer and
Flow Control, its Principles and Application, Vol. 2, edited by G.V. Lachmann,
Pergamon Press, New York, pp. 602–636.
Tournier, S. E., Paduano, J. D., and Pagan, D. (2005), “Flow Analysis and Control in a
Transonic Inlet,” AIAA Paper 2005-4734.
Townsend, A. A. (1956), The Structure of Turbulent Shear Flow, Cambridge University
Press, Cambridge, UK.
Trefethen, L. N., Trefethen, A. E., Reddy, S. C., and Driscoll, T. A. (1993) “Hydrodynamic
Stability without Eigenvalues,” Science, Vol. 261, pp. 578–584.
Tresso R., and Munoz, D. R. (2000), “Homogeneous, Isotropic Flow in Grid Generated
Turbulence,” Journal of Fluids Engineering, Vol. 122, pp. 51–56.
Tseng, Y.-H., and Ferziger, J. H. (2003), “A Ghost-Cell Immersed Boundary Method
for Flow in Complex Geometry,” Journal of Computational Physics, Vol. 192,
pp. 593–623.
Tsien, H. S. (1952), “Servo-Stabilization of Combustion in Rocket Motors,” American
Rocket Society Journal, Vol. 22, pp. 256–263.
Tumin, A. (1996), “Receptivity of Pipe Poiseuille Flow,” Journal of Fluid Mechanics,
Vol. 315, pp. 119–137.
Tumin, A. (2003), “Multimode Decomposition of Spatially Growing Perturbations in
a Two-Dimensional Boundary Layer,” Physics of Fluids, Vol. 15, No. 9,
pp. 2525–2540.
Tumin, A., and Fedorov, A. V. (1984), “Instability Wave Excitation by a Localized Vibrator
in the Boundary Layer,” Journal of Applied Mechanics Technical Physics, Vol. 25,
pp. 867–873.
Tuncer, O., Acharya, S., Cohen, J., and Banaszuk, A. (2005), “Side Air-Jet Modulation for
Control of Heat Release and Pattern Factor,” Combustion Science and Technology,
Vol. 177, No. 7, pp. 1339–1364.
Tung, S., and Kleis, S. (1996), “Initial Streamwise Vorticity Formation in a Two-Stream
Mixing Layer,” Journal of Fluid Mechanics, Vol. 319, No. 1, pp. 251–279.
506 REFERENCES

Tyndall, J. (1881), Sound, D. Appleton, New York, p. 257.


Uhm, J. H., and Acharya, S. (2004), “Control of Combustion Instability with a High-
Momentum Air-Jet,” Combustion and Flame, Vol. 139, pp. 106–125.
Uhm, J. H., and Acharya, S. (2005), “Low-Bandwidth Open-Loop Control of Combustion
Instability,” Combustion and Flame, Vol. 142, pp. 348–363.
Ukeiley, L., Seiner, J., and Ponton, M. (1999), “Azimuthal Structure of an Axisymmetric
Jet Mixing Layer,” ASME FEDSM99-7252.
Ukeiley, L., Cordier, L., Manceau, R., Delville, J., Glauser, M. N., and Bonnet, J. P. (2001),
“Examination of Large Scale Structures in a Turbulent Plane Mixing Layer. Part 2:
Dynamical Systems Model,” Journal of Fluid Mechanics, Vol. 441, pp. 67–108.
Ukeiley, L. S., Ponton, M. K., Seiner, J. S., and Jansen, B. (2004), “Suppression of Pressure
Loads in Cavity Flows,” AIAA Journal, Vol. 42, No. 1, pp. 70–79.
Urbin, G., and Knight, D. (2001), “Large-Eddy Simulation of a Supersonic Boundary
Layer Using an Unstructured Grid,” AIAA Journal, Vol. 39, No. 7, pp. 1288–1295.
Utturkar, Y., Mittal, R., Rampunggoon, P., and Cattafesta, L. (2002), “Sensitivity of
Synthetic Jets to the Design of the Jet Cavity,” AIAA Paper 2002-0124.
Utturkar, Y., Holman, R., Mittal, R., Carroll, B., Sheplak, M., and Cattafesta, L. (2003),
“A Jet Formation Criterion for Synthetic Jet Actuators,” AIAA Paper 2003-0636.
Vaithianathan, T., and Collins, L. R. (2003), “Numerical Approach to Simulating Turbulent
Flow of a Viscoelastic Polymer Solution,” Journal of Computational Physics, Vol. 187,
pp. 1–21.
Vakili, A., Sauerwein, S., and Miller, D. (1999), “Pulsed Injection Applied to Nozzle
Internal Flow Control,” AIAA Paper 99-1002.
Valensi, J., Parigi, H., and Borgel, J. (1942), “Essai d’un Nouveau Dispositif d’Hyper-
sustentation par Soufflage,” Institut Mécanique des Fluides de Marseille, France.
Vasilyev, O. V. (2001), “Computational Constraints on Large Eddy Simulation of
Inhomogeneous Turbulent Complex Geometry Flows,” DNS/LES—Progress and
Challenges, edited by C. Liu, L. Sakell, and T. Beutner, Greyden Press, Columbus, OH,
pp. 93–104.
Vatsa, V. N., and Carpenter, M. H. (2005), “Higher Order Temporal Schemes with Error
Controllers for Unsteady Navier–Stokes Equations,” AIAA Paper 2005-5245.
Vatsa, V. N., and Turkel, E. (2006), “Simulation of Synthetic Jets Using Unsteady Reynolds-
Averaged Navier–Stokes Equations,” AIAA Journal, Vol. 44, No. 2, pp. 217–224.
Verhulst, F. (1996), Nonlinear Differential Equations and Dynamical Systems, 2nd ed.,
Universitext, Springer-Verlag, Berlin.
Viets, H. (1975), “Flip-Flop Jet Nozzle,” AIAA Journal, Vol. 13, No. 10, pp. 1375–1379.
Viets, H., Piatt, M., and Ball, M. (1987), “Boundary Layer Control by Unsteady Vortex
Generation,” Journal of Wind Engineering and Industrial Aerodynamics, Vol. 7,
pp. 135–144.
Visbal, M., and Gaitonde, D. (2001), “Very High-Order Spatially Implicit Schemes for
Computational Acoustics on Curvilinear Meshes,” Journal of Computational Acoustics,
Vol. 9, No. 4, pp. 1259–1286.
Visbal, M. R., and Gordnier, R. E. (2001), “Direct Numerical Simulation of the Interaction
of a Boundary Layer with a Flexible Panel,” AIAA Paper 2001-2721.
Viswanathan, K. (2002), “Analysis of the Two Similarity Components of Turbulent Mixing
Noise,” AIAA Journal, Vol. 40, pp. 1735–1744.
Von Terzi, D. A., Linnick, M. N., Seidel, J., and Fasel, H. F. (2001), “Immersed Boundary
Techniques for High-Order Finite-Difference Methods,” AIAA Paper 2001-2918.
REFERENCES 507

Vorobieff, P., and Rockwell, D. (1996), “Multiple Actuator Control of Vortex Breakdown
on a Pitching Delta-Wing,” AIAA Journal, Vol. 34, pp. 2184–2186.
Wagner, C., Amarouchè, Y., Doyle, P., and Bonn, D. (2003), “Turbulent Drag Reduction of
Polyelectrolyte (DNA) Solutions: Relation with the Elongational Viscosity,”
Europhysics Letters, Vol. 64, pp. 823–829.
Waithe, K. A. (2001), “An Experimental and Computational Investigation of Multiple
Injection Ports in a Convergent-Divergent Nozzle for Fluidic Thrust Vectoring,” M.S.
Thesis, George Washington University, Washington, DC, May.
Walker, S. (1997), “Lessons Learned in the Development of a National Program,” AIAA
Paper 97-334.
Wall, J., Rivir, R., Boxx, I., and Franke, M. (2007), “Effects of Pulsed-D.C. Discharge
Plasma Actuators in a Separated Low Pressure Turbine Boundary Layer,” AIAA Paper
2007-0942.
Walsh, M. (1990), “Riblets,” Viscous Drag Reduction in Boundary Layers, edited by
D. Bushnell and J. Hefner, Vol. 123, pp. 203–261.
Walsh, P. P., and Fletcher, P. (2004), Gas Turbine Performance, 2nd ed., Wiley-Blackwell,
Malden, MA.
Walther, S., Airiau, C., and Bottaro, A. (2001), “Optimal Control of Tollmien–Schlichting
Waves in a Developing Boundary Layer,” Physics of Fluids, Vol. 13, No. 7,
pp. 2087–2096.
Warren, E. S., and Hassan, H. A. (1998), “Transition Closure Model for Predicting
Transition Onset,” Journal of Aircraft, Vol. 35, No. 5, pp. 769–775.
Wei, M., and Freund, J. B. (2006), “A Noise-Controlled Free-Shear Flow,” Journal of Fluid
Mechanics, Vol. 546, pp. 123–152.
Weiberg, J. A., and Dannenberg, R. E. (1954), “Section Characteristics of an NACA 0006
Airfoil with Area Suction near the Leading Edge,” NACA TN 3285.
Weier, T., and Gerbeth, G. (2004), “Control of Separated Flows by Time Periodic Lorentz
Forces,” European Journal of Mechanics, B/Fluids, Vol. 23, pp. 835–849.
Wengle, H., Huppertz, A., Barwolff, G., and Janke, G. (2001), “The Manipulated
Transitional Backward-Facing Step Flow: an Experimental and Direct Numerical
Simulation Investigation,” European Journal of Mechanics B, Vol. 20, pp. 25–46.
Wernz, S., Valsecchi, P., Gross, A., and Fasel, H. F. (2003), “Numerical Investigation of
Turbulent Wall Jets over a Convex Surface,” AIAA Paper 2003-3727.
Wernz, S., Gross, A., and Fasel, H. F. (2005), “Numerical Investigation of Coherent
Structures in Plane and Curved Wall Jets,” AIAA Paper 2005-4911.
Westley, R., and Lilley, G. M. (1952), “An Investigation of the Noise Field from a Small Jet
and Methods for its Reduction,” Rept. No 53. College of Aeronautics, Cranfield, UK.
Widnall, S. E. (1975), “The Structure and Dynamics of Vortex Filaments,” Annual Review
of Fluid Mechanics, Vol. 7, pp. 141–165.
Wiggins, S. (1990), Introduction to Applied Nonlinear Dynamical Systems and Chaos,
Texts in Applied Mathematics, Vol. 2, Springer, Berlin.
Wilcox, D. C. (2000), Turbulence Modeling for CFD, 2nd ed., DCW Industries, La Canada.
Wilkinson, S. (2003), “Oscillating Plasma for Turbulent Boundary Layer Drag Reduction,”
AIAA Paper 2003-1023.
Willems, J. C. (2004), “Deterministic Least Squares Filtering,” Journal of Econometrics,
Vol. 118, No. 1–2, pp. 341–373.
Williams, D. R., and Amato, C. W. (1988), “Unsteady Pulsing of Cylinder Wakes,” AIAA
Paper 88-3532-CP.
508 REFERENCES

Williams, D. R., and Amato, C. W. (1989), “Unsteady Pulsing of Cylinder Wakes,” Frontiers
in Experimental Fluid Mechanics, Vol. 46, Lecture Notes in Engineering, Springer
Verlag, Berlin, pp. 337–364.
Williams, D. R., Acharya, M., Bernhardt, J., and Yang, P. (1991), “The Mechanism of
Flow Control on a Cylinder with the Unsteady Bleed Technique,” AIAA Paper
91-0039.
Williams, D. R., Mansy, H., and Amato, C. (1992), “The Response and Symmetry Properties
of a Cylinder Wake Subjected to Localized Surface Excitation,” Journal of Fluid
Mechanics, Vol. 234, pp. 71–96.
Williams, D., Cornelius, D., Acharya, M., James, D., and Marshall, T. (2006a), “Smart Inlet
Guide Vanes for Active Flow Vectoring in an Axial Compressor,” AIAA Paper 2006-
3517.
Williams, D. R., Cornelius, D., and Rowley, C. W. (2006b), “Supersonic Cavity Response
to Open Loop Forcing,” Proceedings of the First Berlin Conference on Active Flow
Control, Sept.
Williamson, C. H. K., and Roshko, A. (1988), “Vortex Formation in the Wake of an
Oscillating Cylinder,” Journal of Fluids and Structures, Vol. 2, pp. 355–381.
Williamson, C. H. K, Leweke, T., and Miller, G. D. (1998), “Fundamental Instabilities in
Spatially-Developing Wing Wakes and Temporally-Developing Vortex Pairs”, ASME
Paper FEDSM98-4993.
Wiltse, J. M., and Glezer, A. (1993), “Manipulation of Free Shear Flows Using Piezoelectric
Actuators,” Journal of Fluid Mechanics, Vol. 249, pp. 261–285.
Wiltse, J. M., and Glezer, A. (1998), “Direct Excitation of Small-Scale Motions in Free
Shear Flows,” Physics of Fluids, Vol. 10, No. 8, pp. 2026–2036.
Wing, D. J. (1994), “Static Investigation of Two Fluidic Thrust-Vectoring Concepts on a
Two-Dimensional Convergent-Divergent Nozzle,” NASA TM-4574.
Wing, D. J., and Giuliano, V. J. (1997), “Fluidic Thrust Vectoring of an Axisymmetric
Exhaust Nozzle at Static Conditions,” ASME FEDSM97-3228.
Winter, K. G. (1977), “An Outline of the Techniques Available for the Measurement of
Skin Friction in Turbulent Boundary Layers,” Progress in the Aeronautical Sciences,
Vol. 18, pp. 1–57.
Wissink, J. G., and Rodi, W. (2006), “Direct Numerical Simulations of Transitional Flow
in Turbomachinery,” Journal of Turbomachinery, Vol. 128, No. 4, pp. 668–678.
Wo, A. M., Lo, A. C., and Chang, W. C. (2002), “Flow Control via Rotor Trailing-Edge
blowing in Rotor/Stator Axial Compressor,” Journal of Propulsion and Power, Vol. 18,
No. 1, pp. 93–99.
Wolff, S., Brunner, S., and Fottner, L. (2000), “The Use of Hot-Wire Anemometry to
Investigate Unsteady Wake-Induced Boundary-Layer Development on a High-Lift LP
Turbine Cascade,” Journal of Turbomachinery, Vol. 122, No. 4, pp. 644–650.
Wong, W., and Hall, G. (1975), “Suppression of Strong Shock Boundary Layer Interaction
in Supersonic Inlets by Boundary Layer Blowing,” AIAA Paper 75-1209.
Woodcroft, B. (ed.) (1851), The Pneumatics of Hero of Alexandria, Taylor, Walton and
Maberly, London.
Wu, X., and Durbin, P. A. (2001), “Evidence of Longitudinal Vortices Evolved from
Distorted Wakes in a Turbine Passage,” Journal of Fluid Mechanics, Vol. 446,
pp. 199–228.
Wu, J.-Z., Lu, X.-Y., Denny, A. G., Fan, M., and Wu, J.-M. (1998), “Post-Stall Flow Control
on an Airfoil by Local Unsteady Forcing,” Journal of Fluid Mechanics, Vol. 371,
pp. 21–58.
REFERENCES 509

Wygnanski, I. J. (1993), “Interference with Vortex Formation and Control of Fluid Flow to
Reduce Noise and Change Flow Stability,” USA. Patent No. 142730, filed 25 Oct.
1993.
Wygnanski, I. (2004), “The Variables Affecting the Control of Separation by Periodic
Excitation,” AIAA Paper 2004-2505.
Wygnanski, I., and Petersen, R. A. (1987), “Coherent Motion in Excited Free Shear Flows,”
AIAA Journal, Vol. 25, No. 2, pp. 201–213.
Wygnanski, I., Oster, D., and Fiedler, H. (1979), “A Forced Plane Turbulent Mixing Layer:
a Challenge for the Predictor,” Proceedings of the 2nd Symposium on Turbulent Shear
Flow, Imperial College, London.
Xia, H. and Qin, N. (2006), “Comparison of Unsteady Laminar and DES Solutions of
Synthetic Jet Flow,” AIAA Paper 2006-3161.
Yagle, P. J., Miller, D. N., Ginn, K. B., and Hamstra, J. W. (2000), “Demonstration of
Fluidic Throat Skewing for Thrust Vectoring in Structurally Fixed Nozzles,” ASME
2000-GT-0013.
Yagle, P. J., Miller, D. N., Ginn, K. B., and Hamstra, J. W. (2001), “Demonstration of
Fluidic Throat Skewing for Thrust Vectoring,” Transactions of the ASME Journal of
Engineering for Gas Turbines and Power, Vol. 123, No. 3, pp. 502–507.
Yagle, J., Miller, D. N., Bender, E. E., Smith, B. R., and Vermeulen, P. J. (2002), “A
Computational Investigation of Pulsed Ejection,” AIAA Paper 2002-3278.
Yamaleev, N. K., and Carpenter, M. H. (2006), “Quasi-One-Dimensional Model for
Realistic Three-Dimensional Synthetic Jet Actuators,” AIAA Journal, Vol. 44, No. 2,
pp. 208–216.
Yang, V., and Culick, F. E. C. (1986), “Nonlinear Analysis of Pressure Oscillations in
Ramjet Engines,” AIAA Paper 1986-0001.
Yang, V., Kim, S. I., and Culick, F. E. C. (1990), “Triggering of Longitudinal Pressure
Oscillations in Combustion Chambers: I: Nonlinear Gasdynamics,” Combustion
Science and Technology, Vol. 72, pp. 183–214.
Yao, C.-S., Chen, F. J., and Neuhart, D. (2006), “Synthetic Jet Flowfield Database for
Computational Fluid Dynamics Validation,” AIAA Journal, Vol. 44, No. 12,
pp. 3153–3157.
Ye, T., Mittal, R., Udaykumar, H. S., and Shyy W. (1999), “An Accurate Cartesian Grid
Method for Viscous Incompressible Flows with Complex Immersed Boundaries,”
Journal of Computational Physics, Vol. 156, pp. 209–240.
Yehoshua, T., and Seifert, A. (2006a), “Active Boundary Layer Tripping Using Oscilla-
tory Vorticity Generator,” Aerospace Science and Technology, Vol. 10, No. 3,
pp. 175–180.
Yehoshua, T., and Seifert, A. (2006b), “Boundary Condition Effects on the Evolution of a
Train of Vortex Pairs in Still Air,” Aeronautical Journal, Vol. 110, No. 1109, pp. 397–417.
Yi, T., and Gutmark, E. (2007a), “Dynamics of a High Frequency Fuel Actuator and its
Applications for Combustion Instability Control,” Journal of Engineering for Gas
Turbines and Power, Vol. 129, pp. 648–654.
Yi, T., and Gutmark, E. (2007b), “Combustion Instabilities and Control of a Multiswirl
Atmospheric Combustor,” Journal of Engineering for Gas Turbine and Power, Vol. 129,
pp. 31–37.
Yom-Tov, Y., and Seifert, A. (2005), “Multiple Actuators Flow Control over a Glauert Type
Airfoil at Low Reynolds Numbers,” AIAA Paper 2005-5389.
You, D., and Moin, P. (2006), “Large-Eddy Simulation of Flow Separation Over an Airfoil
with Synthetic Jet Control,” CTR Annual Research Briefs, Stanford, CA, pp. 337–346.
510 REFERENCES

You, D., Wang, M., Mittal, R., and Moin, P. (2006a), “Large-Eddy Simulations of
Longitudinal Vortices Embedded in a Turbulent Boundary Layer,” AIAA Journal,
Vol. 44, No. 12, pp. 3032–3039.
You, D., Wang, M., and Moin, P. (2006b), “Large-Eddy Simulation of Flow over a
Wall-Mounted Hump with Separation Control,” AIAA Journal, Vol. 44, No. 11,
pp. 2571–2577.
Young, D. D., Jenkins, S., and Miller, D. N. (2005), “An Investigation of Active Flowfield
Control of Shock/Boundary Layer Interaction,” AIAA Paper 2005-4020.
Yu, K,. and Wilson, K. J. (2002), “Scale-Up Experiments on Liquid-Fueled Active
Combustion Control,” Journal of Propulsion and Power, Vol. 18, pp. 53–60.
Yu, K., Parr, T. P., Wilson, K. J., Schadow, K. C., and Gutmark, E. (1996), “Active Control
of Liquid Fueled Combustors Using Periodic Vortex-Droplet Interactions,” Proceedings
of the Combustion Institute, Vol. 26, pp. 2843–2850.
Yuan, L. L., Street, R. L., and Ferziger, J. H. (1999), “Large-Eddy Simulations of a Round
Jet in Crossflow,” Journal of Fluid Mechanics, Vol. 379, pp. 71–104.
Yuan, C. C. L., Krstic, M., and Bewley, T. R. (2004), “Active Control of Jet Mixing,” IEEE
Control Theory and Applications, IET, Vol. 151, No. 6, pp. 763–772.
Yurchenko, N., and Rivir, R. B. (2000), “Improvement of the Turbine Blade Performance
Based on the Flow Instability and Receptivity Analysis,” 8th International
Symposium on Transport Phenomena and Dynamics of Rotating Machinery, Vol. 1,
pp. 300–306.
Zacharos, A., and Kontis, K. (2004), “Numerical Studies on Active Flow Circulation
Controlled Flap Concept for Aeronautical Applications,” AIAA Paper 2004-2314.
Zaman, K. B. M. Q. (1999), “Spreading Characteristics of Compressible Jets from Nozzles
of Various Geometries,” Journal of Fluid Mechanics, Vol. 383, pp. 197–228.
Zaman, K., and Culley, D. (2006), “A Study of Stall Control over an Airfoil Using Synthetic
Jets,” AIAA Paper 2006-0098.
Zaman, K. B. M. Q., and Raman, G. (1997), “Reversal in Spreading of a Tabbed Circular
Jet Under Controlled Excitation,” Physics of Fluids, Vol. 9, No. 12, pp. 3733–3741.
Zaman, K. B. M. Q., Bar-Sever, A., and Mangalam, S. M. (1987), “Effect of Acoustic
Excitation on the Flow over a Low Re Airfoil,” Journal of Fluid Mechanics, Vol. 182,
pp. 127–148.
Zaman, K. B. M. Q., Reeder, M. F., and Samimy, M. (1994), “Control of an Axisymmetric
Jet Using Vortex Generators,” Physics of Fluids, Vol. 6, No. 2, pp. 778–793.
Zang, T. A. (1991), “On the Rotation and Skew-Symmetric Forms for Incompressible Flow
Simulations,” Applied Numerical Mathematics, Vol. 7, pp. 27–40.
Zhang, H. L., Bachman, C. R., and Fasel, H. F. (2000), “Application of a New Methodology
for Simulations of Complex Turbulent Flows,” AIAA Paper 2000-2535.
Zhang, X. (2003), “The Evolution of Co-Rotating Vortices in a Canonical Boundary Layer
with Inclined Jets,” Physics of Fluids, Vol. 15, No. 12, pp. 3693–3702.
Zhang, X. F., and Hodson, H. (2005), “Combined Effects of Surface Trips and Unsteady
Wakes on the Boundary Layer Development of an Ultra-High-Lift LP Turbine Blade,”
Journal of Turbomachinery, Vol. 127, No. 3, pp. 479–488.
Zhao, H., and Bau, H. H. (2006) “Limitations of Linear Control of Thermal Convection in
a Porous Medium,” Physics of Fluids, Vol. 18, pp. 074109–12.
Zhao, Y., Collins, E. Jr., Alvi, F., and Dores, D. (2005), “Design and Implementation of
Feedback Flow Control for Counterflow Thrust Vectoring,” Journal of Propulsion and
Power, Vol. 21, No. 5, pp. 815–821.
REFERENCES 511

Zhou, M. D., and Wygnanski, I. (2001), “The Response of a Mixing Layer Formed between
Parallel Streams to a Concomitant Excitation at Two Frequencies,” Journal of Fluid
Mechanics, Vol. 441, pp. 139–168.
Zhou, K., Doyle, J. C., and Glover, K. (1996), Robust and Optimal Control, Prentice-Hall,
Upper-Saddle River, NJ.
Zhuang, N., Alvi, F. S., Alkislar, M. B., and Shih, C. (2006), “Supersonic Cavity Flows and
their Control,” AIAA Journal, Vol. 44, No. 9, pp. 2118–2128.
Zinn, B. T. (2005), “Smart Combustors – Just Around the Corner,” ASME Turbo Expo
2005: Power for Land, Sea, and Air (GT2005), June 6–9, 2005, Reno, Nevada, USA,
Vol. 2: Turbo Expo 2005, Paper no. GT2005-69138, pp. 1011–1031.
Zinn, B. T., and Lieuwen, T. (2005), “Combustion Instabilities: Basic Concepts,”
Combustion Instabilities in Gas Turbine Engines, edited by T. Lieuwen and V. Yang,
Vol. 210, Progress in Astronautics and Aeronautics, pp. 3–26.
Zinn, B. T., and Neumeier, Y. (1997), “An Overview of Active Control of Combustion
Instabilities,” AIAA Paper 97-0461.
Zoppellari, E., and Juve, D. (1997), “Reductions of Jet Noise by Water Injection,” AIAA
Paper 1997-1622-CP.
Zuccher, S., Tumin, A., and Reshotko, E. (2006), “Parabolic Approach to Optimal
Perturbations in Compressible Boundary Layers,” Journal of Fluid Mechanics, Vol.
556, pp. 189–216.
INDEX

Index Terms Links

Acoustic control of jets 365


Acoustic feedback 334
Acoustically compact surfaces 357
Active control 331
Active flow control (AFC) 10 69
actuator 26
aerodynamics 15
aerodynamics, actuator technology 16
aerodynamics, oscillating-flap actuators 16
applications 405
helicopter blade-vortex interactions 406
high-speed impulsive noise 418
oscillatory jets 424
XV-15 tiltrotor hover download reduction 431
closed-loop 241
computational fluid dynamics 18
conditional sampling 25
decomposition 23
delay boundary-layer transition 17
design variable, active flow control 278
flight conditions 49

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Active flow control (AFC) (Cont.)


free shear layer 27
high-lift systems 233
implementation in jet engines 319
jets 34
large-amplitude perturbations 34
small-amplitude perturbations 34
Navier-Stokes equations 22
numerical simulations 302
controlled flow 306
uncontrolled flow 303
optimal excitation frequency 250
passive flow technique 17
Prandtl’s boundary-layer theory 24
scaling 253
separation concept 24
separation of control 37
superlayer 24
triple decomposition 25
turbulence 23
turbulence production mechanism 17
variable-interval time-averaging 25
wake vortex control 50
zone averaging 25
Active separation control (ASC) 232
high-lift systems 233

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Actuators 26 61 150 333


calibration of 27
fundamental principles of 150
history of 150
sensors vs 149
technology 16
technology, synthetic jet 16
types 153
fluidic 153
non-thermal plasma 164
Adaptive control 148
Adaptive Flow Control Vehicle Integrated
Technologies (ADVINT) 238
Boeing tilt-wing 238
Adjoint-based optimization 83
ordinary-differential equations 84
partial-differential equations 84
receptivity 83
Adjoint-based theoretical flow control 75
Adjoints
linear control problem 88
linear state-space control systems 86
Riccati equation 88
Adverse pressure gradient boundary
layers 103
ADVINT. See Adaptive Flow Control Vehicle
Integrated Technologies.

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Aeroacoustics 353
acoustic control of jets 365
circulation control 366
leading edge noise 359
separated flows 365
sound generation by flow 354
synthetic jets 368
trailing edge noise 363
Aerodynamics
active flow control 15
oscillating-flap actuators 16
Aeromechanics 4
Aeronautics 3
finite-span wing effects 3
AFC. See active flow control.
Ahmed car body 246
Air-breathing propulsion flowpath
applications 373
future of 375
inlet aperture 378
inlet duct 383
nozzle 391
Airfoils 9
high lift coefficients 9
leading-edge curvature 39
Stratford pressure recovery 9
Algebraic Riccati Equation 129
Amplified disturbances 127
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Approximately controllable systems 65


Archimedean screw 21
Area rule 127
ASC. See active separation control.
Attachment 45
Attenuated disturbances 127
Axis flow control 52

Balanced truncation 141


Balancing transformation 141
Bandwidth 64
Bernoulli effect 3 52
Bifurcations 144
global 145
local 145
simple 144
Blade mounted trailing edge flap 407
Blade vortex interactions 359
Blades, simulation of 287
open-loop control 293
uncontrolled flow 289
Blasius boundary layer 101 102
BLC. See boundary layer control.
Blended RANS-LES CFD model 200
Blown flap technique 5
Bluff bodies 241

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Bluff-body control 41
Bode plot 124
Bode’s integral function 127
Boeing tilt-wing 238
Boundary conditions 182 190 196 228
Boundary layer
externally blown flaps 8
laminary 8
turbulent 8
Boundary layer control (BLC) 4 22 232
aeromechanics 4
delay separation 4
hydromechanics 4
lift enhancement 5
Boundary layer theory 3 4 24
Boundary methods 202
Braid wake instability 55
BVI. See helicopter blade-vortex
interactions

Canonical flows 12
closed-loop control 12
open-loop control 12
CC. See circulation control.
CF. See crossflow.
CFD. See computational fluid dynamics.

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

CIACS. See combustion instability active


control system.
Circular cylinder 41
bluff-body control 41
Circular synthetic jet into turbulent crossflow
boundary layer 216
Circulation control (CC) 211 366
CLAFC. See closed-loop active flow
control.
Classical closed-loop control 121
gain margins 125
phase margin 126
proportional-integral-derivative
feedback 121
robustness of 126
sensitivity function 126
stability 125
transfer functions 123
Classical control 63
Closed-loop active flow control
(CLAFC) 241
Closed-loop flow control 12 60 227 334
examples 133
open loop vs 318
opposition control 227
ordinary differential equations 133
sensors and 169

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Closed-loop stability 125


Nyquist plot 125
Closed-loop system 67
architecture 116
controllability 119
feedback 116
frequency domain 121
Lyapunov equations 120
multi-input, multi-output 117
observability 119
state space 121
classical 121
feedback loop 67
Closure problem 23
Coanda cylinder 38
Combustion
actuators 166
control 321
acoustic feedback 334
active control system 335
actuators 333
closed-loop 334
hysteresis 334
lean blowout 324 328 341
open-loop 333
secondary fuel injection 334

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Combustion (Cont.)
instability 324
numerical studies of 338
passive control of 330
physics of 324
Combustion instability active control system
(CIACS) 335
experiments 332
Compensator 61
Complementary sensitivity transfer
function 64
Compressible flat-plate zero boundary
layers 103
Computational effort 228
Computational fluid dynamics (CFD) 18
advanced studies 219
control theory 227
neural networks 226
parameter optimization 225
plasma actuators 164 220
polymers 222
methodologies 178
direct numerical simulation 178
Large-eddy simulation 187
Reynolds-averaged Navier-Stokes 195
models 200
(CFD)
blended RANS-LES 200
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Computational fluid dynamics (CFD) (Cont.)


immersed boundary methods 202
models, reduced-order 202
Navier-Stokes equation 18
numerical simulations 18
turbulent flow 18
validation 211
circular synthetic jet into turbulent
crossflow boundary layer 216
circulation control 211
2-D flow control for separation control 217
2-D synthetic jet into quiescent air 214
Conditional sampling 25
Continuous forcing approach 203
Control systems 170
Control theory 227
boundary conditions 228
closed-loop 227
computational effort 228
Control-affine form 143
Controllability 65 119
Gramian 120
in finite time 65
Controlled flow 306
pulsed vortex-generator jets 161 309
steady vortex-generator jets 306
Controlled output 117
Controller 61
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Controlling flow instabilities 12


Convergence 118
Cost benefit, low-pressure turbines 317
Crossflow (CF) 104
Curved plate geometry 313

DBD. See dielectric barrier discharges.


Decompositions 23 80
Reynolds stresses 23
Delay boundary-layer transition 17
Delay separation 4
Delta wings 246
Detectable systems 65
Developing background flows 99
adverse pressure gradient boundary
layers 103
Blasius boundary layer 101 102
compressible flat-plate zero boundary
layers 103
crossflow 104
Gortler vortices 102
incompressible flow limit 105
parabolized stability equations 99
Dielectric barrier discharges (DBD) 281
Dimples 265
Direct MEMS sensors 174

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Direct numerical simulation 178


boundary conditions 182
equations 178
Navier-Stokes equations 178
Kolmogorov dissipation length
scale 183
numerical considerations 183
Discrete forcing approach 205
finite volume formulation 208
Fourier pseudospectral method 207
ghost cell method 207
Distributed roughness elements (DRE) 90
Distributed suction 7
laminar boundary layers 8
turbulent boundary layers 8
Disturbances 127
amplified 127
attenuated 127
Drag increase 245
fluidic spoilers 245
wake manipulation 245
Drag reduction 222 241
bluff bodies 241
streamlined bodies 241
DRE. See distributed roughness elements.
DTN. See dual throat nozzle.
Dual throat nozzle (DTN) 397
Dynamic feedback 69
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Dynamic stall control 47


Dynamic vortex perturbations 55
braid wake instability 55
long-wave instability 55
short-wave instability 55

Eigenmode control 109


Electro-hydrodynamic actuation 44
Equations 178 187 195
Large eddy simulation 187
Navier-Stokes 178
Reynolds-averaged Navier-Stokes 195
Euler 3
Externally blown flaps 8
Extremum seeking 148

Feedback 66 116
dynamics 116
full-state 129
observer-based 131
sensitivity reduction 117
Feedback control 60
Feedback loop 67
Feedforward 66

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

FENE. See Finitely Extensible Nonlinear


Elastic model.
Finite volume formulation 208
Finite-span wing effects 3
Finitely Extensible Nonlinear Elastic model
(FENE) 224
Peterlin approximation 224
Fixed wing airfoil application 231
active separation control 232
drag increase 245
drag reduction 241
moment generation 243
vortex flows 246
Fixed wing separation control, boundary-layer
control 232
Flight conditions, active flow control and 49
Flight testing 436
Flow control 64 306
active 12 69
adjoint-based optimization 83
adjoint-based theoretical approach 75
aeroacoustics 353
boundaries of 71
classification of 70
closed-loop system 67
definition of 65
dynamic feedback 69
feedback 66
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Flow control (Cont.)


feedforward 66
flow-field physics 20
history of 1
aeronautics 3
Bernoulli 3
Euler 3
Hero 2
Leonardo da Vinci 2
Prandtl 3
instabilities 12
loop transfer function 68
modern 10
open-loop 67
passive 9 69 263 330
quasi-steady feedback 69
rotorcraft applications 403
terminology 59
actuators 61
approximately controllable 65
bandwidth 64
classical control 63
closed loop 60
compensator 61
complementary sensitivity transfer
function 64
controllable in finite time 65
controller 61
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Flow control (Cont.)


detectable 65
feedback 60
frequency domain 62
frequency response 62
gain-scheduled 64
loop crossover frequency 64
loop transfer function 64
low-order model 63
model reduction 63
modeling 61
modern control 63
observability 65
open loop 60
partial differential equations 63
reachable 65
reduced-order model 63
robust control 63
sensitivity transfer function 64
sensors 61
SISO controller 64
stability 65
stabilizable 65
state variables 62
state-space form 62
system identification techniques 62
time domain 62

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Flow control (Cont.)


transfer function 62
unsteady 66
Flows, fully turbulent 11
Fluid thrust
counterflow vectoring 392
dual throat nozzle 397
flow control 391
pulsed actuators 159 400
reduced injector flow rate 400
shock vector 394
virtual aero surface shaping 394
Fluidic actuators 153
combustion 166
Hartmann tube 162
piezoelectric flaps 158
powered resonance tubes 162
scaling of 163
SparkJet 167
synthetic jets 153
Fluidic spoilers 245
Fourier pseudospectral method 207
Free shear layer 27
three-dimensional
development 31
perturbations 32
two-dimensional perturbations 28
Free turbulence 354
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Free-stream turbulence (FST) 260


Frequency domain 62 121
Frequency response 62 124
Bode plot 124
FST. See free-stream turbulence.
Fuel injection 334
Full-state feedback 129
linear quadratic regulator problem 129
Functional modeling 189

Gain margins 64 125


Gain scheduling 64 147
Galerkin projection 138
Ghost cell method 207
Global bifurcations 145
Gortler instability 39
Gortler vortices 102
Gramian controllability 120
Gramian observability 120

Hartmann tube 162


Helicopter blade-vortex interactions (BVI) 406
plain flap drag reduction 410
steady blowing 411

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Helicopter blade-vortex interactions (BVI) (Cont.)


suction 411
trailing-edge flap 407 416
Hero 2
High lift
airfoil 9
coefficients 9
separation of control 37
systems 233
Adaptive Flow Control Vehicle Integrated
Technologies 238
NASA-TAU simplified 233
High-speed impulsive noise 418
shock waves 418
Hydromechanics 4
Hysteresis control 334

ILES. See implicit modeling.


Immersed boundary methods 202
continuous forcing approach 203
discrete forcing approach 205
incompressible flow problems 210
Implementation, sensors and 170
Implicit modeling (ILES) 189
Incompressible dynamic stall
control 47

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Incompressible flow 79
limit 105
problems 210
Indirect MEMS sensors 173
Initial value problem 77
Inlet aperture 378
shock/boundary layer flow control 378
Inlet duct 383
separation control 383
vortex flow control 383
Instability theory 73
applications of 90
developing background flows 99
distributed roughness elements 90
laminar flow control 91
no flow-natural convection 92
non-parallel background flows 107
parallel background flows 93
three-dimensional flows 112
Tollmien-Schlichting waves 90
ultrasonic absorptive coatings 90
decomposition 80
elements of 76
incompressible flow 79
initial value problem 77
linear operator A 79
linearized Navier-Stokes equations 73
turbulent flows 82
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Instability, wave 55
Internally blown flap technique 5

Jacobian matrix 143


Jets
control of 34 365
small-amplitude perturbations 34

Kalman filter 63
Kolmogorov dissipation length scale 183

Laminar flow 23
Laminar flow control (LFC) 91
Laminary boundary layers 8
Laplace transform 123
Large-amplitude perturbations 34
Large-eddy simulation (LES) 187 338
boundary conditions 190
equations 187
functional modeling 189
implicit modeling 189
limitations 194
numerical considerations 191

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Large-eddy simulation (LES) (Cont.)


Smagorinsky model 190
structural modeling 189
turbulence modeling 187
LBO. See lean blowout.
Leading edge noise 359
blade vortex interactions 359
rotor stator interactions 361
Lean blowout (LBO), control of 341
lean prevaporized premixed combustion 341
numerical studies 348
physics of 328
studies 342
Lean prevaporized premixed (LPP)
combustion 341
LEM. See lumped element modeling.
Leonardo da Vinci 2
LES-RANS model 200
LES. See large-eddy simulation.
Leveraging flow instabilities 10
LFC. See laminar flow control.
Lift enhancement 5
blown flap technique 5
L1M blade 292 298
Linear control problem 88
Linear operator A, dimensionality of 79
Linear Quadratic Gaussian (LQG) 63
regulator 131
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Linear Quadratic Regulator (LQR) 63 88


problem 129
algebraic Riccati Equation 129
Linear state-space control system 86
Linearized Navier-Stokes equations (LNSE) 73
LNSE. See linearized Navier-Stokes equations.
Local bifurcations 145
Long-wave instability 55
Loop crossover frequency 64
Loop transfer function 64 68
Low incidence control 243
Low-order model 63
Low-pressure turbines (LPTs) 259
active flow control, design variable 278
cost/gain estimate 317
experiments 263
free-stream turbulence 260
open vs closed loop control 318
passive actuators 263
plasma actuators 164 281
pulsed jet actuation 271 299 306
separation control 263 281 293 306
simulations 286
simulations, blades 287
simulations, model geometry 302
steady jet actuation 266 306
vortex-generator jets 266 299 306
Low-Reynolds-number airfoils 9
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

LPP. See lean prevaporized premixed.


LPTs. See low-pressure turbines.
LQG. See linear-quadratic-Guassian.
LQR. See linear-quadratic regulator.
Lumped element modeling (LEM) 155
Lyapunov equations 120

Magneto-hydrodynamic acutation 44
MEMs. See microelectromechanical systems.
Micro aerial vehicles 49
Microelectromechanical Systems (MEMS) 149 170
listing of 171
sensors
direct 174
indirect 173
unsteady pressure sensors 172
wall shear stress 173
MIMO. See multi-input, multi-output
systems.
Minimum-variance optimal estimation 63
Kalman filter 63
MinMax control 132
Model reduction 63 136
Galerkin projection 138
proper orthogonal decomposition 139

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Modeling
flow control and 61
functional 189
low-order 63
reduced order 63
Smagorinsky 190
structural 189
Modern flow control 10 63
active 12
canonical flows 12
evolution of 11
instabilities 10
linear-quadratic regulator 63
linear-quadratic-Gaussian 63
minimum-variance optimal estimation 63
unsteady actuation 11
Moment generation 243
low incidence control 243
reattachment 243
separating attached flow 243
Moving surfaces, turbulence and 356
Multi-input, multi-output (MIMO) systems 117
controlled output 117
convergence 118
Multiple equilibria 142
control-affine form 143
equilibrium points 142
Jacobian matrix 143
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Nano air vehicles 49


NASA-TAU simplified high-lift system 233
Natural convection 92
Navier-Stokes equation 18 22 178
proper orthogonal decomposition 18
Reynolds 23
Neural networks 226
No flow—natural convection 92
Non-parallel background flows 107
eigenmode control 109
Non-thermal plasma actuators 164
single dielectric barrier discharge 164
Nonlinear oscillations 146
Nonlinear systems 142
control of 147
gain scheduling 147
receding horizon control 147
extremum seeking 148
multiple equilibria 142
oscillations 146
periodic orbits 144
Poincare map 144
simple bifurcations 144
Noria 21
Nozzles 391
fluidic thrust vectoring flow control 391
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Numerical simulation 18
Nyquist plot 125

Observability 65 119
Gramian 120
Observer-based feedback 131
Linear Quadratic Gaussian regulator 131
separation principle 131
ODEs. See ordinary-differential equations.
Open-loop flow control 13 60 67 271
293 306 333
closed loop vs 318
law 147
L1M blade 281
Pack B blade 263
uncontrolled 68
Opposition control 227
Optimal control theory 227
robust 228
Optimal excitation frequency 250
Ordinary-differential equations
(ODEs) 84 133
Oscillating-flap acuators 16
Oscillations 146
probability density function 147
Oscillatory jets, stall/post-stall lift 424

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Pack B blade 289


Parabolized stability equations (PSE) 99
Parallel background flows 93
Parameter optimization 225
Partial-differential equations (PDE) 63 84
Particle image velocimetry (PIV) data 17 276
Passive actuators 263
passive flow control 265
Passive flow control (PFC) 9 69 263 330
dimples 265
high-lift airfoil 9
low-Reynolds-number airfoils 9
technique 17
PDE. See partial-differential equations.
PDF. See probability density function.
Periodic orbits 144
Perturbations 28 32
Peterlin approximation 224
PFC. See passive flow control.
Phase margin 64 126
PID feedback 121
Piezoelectric flaps 158
PIV. See particle image velocimetry.
Plain flap drag reduction 410
Plasma actuators 45 164 220 281
dielectric barrier discharges 281
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

POD. See proper orthogonal decomposition.


Poincare map 144
Pole-placement theorem 130
Polymeric drag reduction 222
Finitely Extensible Nonlinear Elastic model 224
Post-stall enhancement 424
Powered resonance tubes (PRT) 162
Prandtl
boundary layer 4 22
control 22
theory 3 24
Probability density function (PDF) 147
Process noise 130
Proper orthogonal decomposition (POD) 18 139 310
particle-image-velocimetry data 18
Proportional-integral-derivative (PID)
feedback 121
PRT. See powered resonance tubes.
PSE. See parabolized stability equations.
Pulsed actuators 159 400
Pulsed jet actuation 271 299 306
experiments 271
particle image velocimetry data 276
simulations 299 306
Pulsed jets 159
Pulsed microjets 160

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Pulsed vortex-generator jets 161 271 299 306


experiments 271
simulations, curved plate geometry 313
simulations, flat plate geometry 306
simulations, L1M blade 299

Quasi-steady feedback 69
Quiescent flow 23

RANS-LES model 200


RANS. See Reynolds-averaged Navier-Stokes.
Receding horizon control 147
open-loop control law 147
Receptivity 83
Reduced injector flow rate 400
Reduced-order model 63 202
Reynold’s averaging 23
Reynolds numbers 49
Reynolds stresses 23
Reynolds-averaged Navier-Stokes (RANS)
boundary conditions 196
equations 195
limitations of 199
methods 23

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Reynolds-averaged Navier-Stokes (RANS) (Cont.)


numerical considerations 198
turbulence modeling 195
Riccati equation 88 129
Robust control 63
theory 228
Robustness 126
Rotor stator interactions 361
Rotorcraft applications flow control 403

SBLI. See shock/boundary layer interaction.


Scaling 253
SDBD. See single dielectic barrier discharge.
Sensitivity function 64 126
area rule 127
Bode’s integral function 127
disturbances 127
Sensitivity reduction 117
Sensitivity transfer function 64
Sensor noise 130
Sensors 61 168
actuators vs 149
basics of 168
closed-loop flow control
strategies 169
microelectromechanical 170

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Sensors (Cont.)
traits
control systems 170
implementation 170
non-intrusive nature 169
spatial resolution 169
temporal resolution 169
Separated flows 243 365
Separation control 37 263 281 293
306 383 397
attachment 45
electro-hydrodynamic actuation 44
flight conditions 49
high lift 37
incompressible dynamic stall control 47
magneto-hydrodynamic actuation 44
micro aerial vehicles 49
nano air vehicles 49
plasma actuators 45 164
Reynolds numbers 49
steady jet actuation 266 306
streamwise curvature 38
unsteady 45
Separation principle 24 131
Shock vector control (SVC) 394
Shock waves 418
Shock/boundary layer flow control 378
Shock/boundary layer interaction (SBLI) 378
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

Short-wave instability 55
Simple bifurcations 144
Single dielectric barrier discharge (SDBD) 164
Sinuous flow 12
SISO controller 64
complementary sensitivity transfer
function 64
gain margin 64
loop crossover frequency 64
loop transfer function 64
phase margin 64
sensitivity transfer function 64
Smagorinsky model 190
Small-amplitude perturbations 34
Sound generation by flow 354
acoustically compact surfaces 357
large surfaces 358
turbulence 354
Sparkjet actuators 167
Spatial resolution 169
SSL. See suction surface length.
Stability 65
Stall enhancement 424
Stall phenomena 7
State estimation 129
pole-placement theorem 130
process noise 130
sensor noise 130
This page has been reformatted by Knovel to provide easier
navigation.
Index Terms Links

State space 121


State variables 62
State-space form 62
Steady blowing 411
Steady jet actuation 266 306
Steady vortex-generator jets 266 306
Stratford pressure recovery 9
Streamlined bodies, drag reduction in 241
Streamwise curvature 38
airfoil leading-edge 39
circular cylinder 41
Coanda cylinder 38
Gortler instability 39
vortex shedding frequency 44
Structural modeling 189
Suction surface length (SSL) 303
Suction 411
Superlayer, concept of 24
SVC. See shock vector control.
Synthetic jets 153 368
pulsed jets 159
technology 16
vortex generating jets 161
Zero-net mass-flux 153
System identification techniques 62
System dynamics 116

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Temporal resolution 169


Three-dimensional flows 31 112
Throat skewing 394
Tiltrotor hover download reduction 431
flight tests 436
numerical simulations 431
wind tunnel experiments 436
zero mass flux forcing 437
Time domain 62
Tollmien-Schlichting (TS) waves 90
Trailing-edge flap, transpiration patch 416
Trailing-edge noise 363
Transfer functions 62 123
frequency response 124
Laplace transform 123
Transpiration patch 416
Triple decomposition 25
TS. See Tollmien-Schlichting.
Tubes
Hartmann 162
power resonance 162
Turbomachinery applications 259
Turbulence 354
modeling 187 195
moving surfaces 356

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Turbulence (Cont.)
problem 23
production mechanism 17
Turbulent boundary layers 8
Turbulent flow 18 23 82
Two-dimensional flow control for separation
control 217
Two-dimensional perturbations 28
Two-dimensional synthetic jet into
quiescent air 214

UAC. See ultrasonic absorptive coatings.


Ultrasonic absorptive coatings (UAC) 90
Uncontrolled flow 289 293 303
L1M blade 292 298
Pack B blade 289
suction surface length 303
Uncontrolled system 68
Unsteady actuation 11
fully turbulent flows 11
Unsteady flow control 66
Unsteady pressure sensors 172
Unsteady separation 45

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Variable-interval time averaging (VITA) 25


Vehicles
micro aerial 49
nano air 49
VGJs. See vortex generator jets.
Virtual aero surface shaping
separation control 397
throat skewing 394
VITA. See variable-interval time averaging.
Vortex flows 246
Ahmed car body 246
control 50 383
delta wings 246
fuselage 246
Vortex generator jets (VGJs) 161 266 299 306
pulsed 271 299 306
steady 266 306
Vortex perturbation 50
Vortex shedding frequency (VSF) 44
VSF. See vortex shedding frequency.

Wake manipulation 245


Wake vortex control 50
axis flow control 52

This page has been reformatted by Knovel to provide easier


navigation.
Index Terms Links

Wake vortex control (Cont.)


basics 50
Bernoulli effect 52
dynamic perturbations 55
Wall shear stress 173
direct MEMS sensors 174
indirect MEMS sensors 173
Wind tunnel 436

Zero mass flux forcing (ZMFF) 437


Zero-net mass flux actuators (ZNMF) 153
lumped element modeling 155
ZMFF. See zero mass flux forcing.
ZNMF. See zero-net mass flux actuators.
Zone averaging 25

This page has been reformatted by Knovel to provide easier


navigation.
Supporting Materials

Many of the topics introduced in this book are discussed in more detail in
other AIAA publications. For a complete listing of titles in the AIAA Progress
Series, as well as other AIAA publications, please visit www.aiaa.org.
AIAA is committed to devoting resources to the education of both practicing
and future aerospace professionals. In 1996, the AIAA Foundation was founded.
Its programs enhance scientific literacy and advance the arts and sciences of
aerospace. For more information, please visit www.aiaafoundation.org.

You might also like