improved system idenification - oklahoma state … · improved system identification for...

179
IMPROVED SYSTEM IDENTIFICATION FOR AEROSERVOELASTIC PREDICTIONS By CHARLES ROBERT O’NEILL Bachelor of Science Oklahoma State University Stillwater, Oklahoma 2001 Submitted to the Faculty of the Graduate College of Oklahoma State University in partial fulfillment of the requirements for the Degree of MASTER OF SCIENCE August, 2003

Upload: dangxuyen

Post on 05-Jul-2018

222 views

Category:

Documents


0 download

TRANSCRIPT

IMPROVED SYSTEM IDENTIFICATION

FOR AEROSERVOELASTIC

PREDICTIONS

By

CHARLES ROBERT O’NEILL

Bachelor of Science

Oklahoma State University

Stillwater, Oklahoma

2001

Submitted to the Faculty of the Graduate College of Oklahoma State University

in partial fulfillment of the requirements for

the Degree of MASTER OF SCIENCE

August, 2003

ii

IMPROVED SYSTEM IDENTIFICATION

FOR AEROSERVOELASTIC

PREDICTIONS

Thesis Approved:

________________________________________________

Thesis Advisor

________________________________________________

________________________________________________

________________________________________________ Dean of the Graduate College

iii

ACKNOWLEDGEMENTS

For your patience, understanding, encouragement, support, lessons and love, I offer my

deepest gratitude. Thank you.

iv

TABLE OF CONTENTS

Chapter Page

1 INTRODUCTION ........................................................................................................... 1

1.1 Background......................................................................................................... 1 1.2 Current Status ..................................................................................................... 3 1.3 Simulation Overview .......................................................................................... 4

1.3.1 Structure.......................................................................................................... 4 1.3.2 Aerodynamics ................................................................................................. 6 1.3.3 System Identification ...................................................................................... 8

1.4 Objectives ........................................................................................................... 9

2 LITERATURE REVIEW .............................................................................................. 10

2.1 Unsteady Aerodynamics ................................................................................... 10 2.2 Aerodynamic System Models........................................................................... 14

2.2.1 Indicial Methods ........................................................................................... 14 2.2.2 ARMA........................................................................................................... 15 2.2.3 Nonlinear....................................................................................................... 17

2.3 System Identification Methodologies ............................................................... 19 2.4 Input Training Signals ...................................................................................... 23 2.5 Model Quality ................................................................................................... 25 2.6 Aerodynamic and Structural System Representations...................................... 26

2.6.1 Structural Model ........................................................................................... 26 2.6.2 Aerodynamic Model ..................................................................................... 27 2.6.3 Aeroelastic Model......................................................................................... 28 2.6.4 Aeroservoelasticity Model ............................................................................ 30

3 METHODOLOGY ........................................................................................................ 33

3.1 Aerodynamics System Model........................................................................... 33 3.1.1 CFD Solver ................................................................................................... 33 3.1.2 Aerodynamic Specific Requirements ........................................................... 34 3.1.3 Objective Function........................................................................................ 35 3.1.4 Description Function Selection..................................................................... 37 3.1.5 ARMA Realizations and Canonical Forms................................................... 38 3.1.6 ARMA Model Transfer Function ................................................................. 42 3.1.7 ARMA Pathology ......................................................................................... 42

3.2 Training Method ............................................................................................... 44 3.2.1 System Identification Data Flow .................................................................. 44

v

3.2.2 SVD Data Flow............................................................................................. 45 3.2.3 Time Scales and Aero-Structural Integration ............................................... 46 3.2.4 Training Data Redundancy ........................................................................... 48 3.2.5 Serial Training .............................................................................................. 49 3.2.6 Parallel Training............................................................................................ 49 3.2.7 Model Splicing.............................................................................................. 50

3.3 Excitation Signals ............................................................................................. 53 3.3.1 Criteria .......................................................................................................... 54 3.3.2 3211 Multistep .............................................................................................. 55 3.3.3 Variable Amplitude Multistep ...................................................................... 58 3.3.4 Chirp ............................................................................................................. 59 3.3.5 DC-Chirp....................................................................................................... 62 3.3.6 Fresnel Chirp................................................................................................. 65 3.3.7 Schroeder Sweep........................................................................................... 68 3.3.8 Noise Training Signal ................................................................................... 72 3.3.9 Envelopes...................................................................................................... 74 3.3.10 Superposition of Multiple Signals ............................................................ 75 3.3.11 Purposefully Added Noise ........................................................................ 76 3.3.12 Motion Specification................................................................................. 80

3.4 Model Performance Evaluation Criteria ........................................................... 82 3.4.1 Chi Squared................................................................................................... 83 3.4.2 Force Prediction Root Mean Square ............................................................. 84 3.4.3 Partial Autocorrelation.................................................................................. 85 3.4.4 Coupled Aero-Structural Properties.............................................................. 86

3.5 Preliminary Testcases ....................................................................................... 90 3.5.1 Zero Order Force Function ........................................................................... 90 3.5.2 Second Order Force Function ....................................................................... 95

4 RESULTS .................................................................................................................... 102

4.1 Aerodynamic System Identification Training Method ................................... 102 4.2 Single Degree of Freedom Divergence........................................................... 103

4.2.1 Mach 2.0 ..................................................................................................... 106 4.2.2 Mach 0.6 ..................................................................................................... 111

4.3 AGARD 445.6 ................................................................................................ 118 4.3.1 Flutter Boundary ......................................................................................... 119 4.3.2 Sensitivity Studies....................................................................................... 127

4.4 Panel Flutter .................................................................................................... 130 4.4.1 Serial Chirp Training .................................................................................. 132 4.4.2 Parallel Chirp Training ............................................................................... 134 4.4.3 Free Response Aeroelastic Boundary Validation ....................................... 135

4.5 Wing/Flap Control .......................................................................................... 136 4.5.1 Aerodynamic and Structural Representations............................................. 137 4.5.2 Training....................................................................................................... 138 4.5.3 Controls....................................................................................................... 139

5 CONCLUSIONS AND RECOMMENDATIONS ...................................................... 143

vi

5.1 Conclusions..................................................................................................... 143 5.2 System Identification Recommendations ....................................................... 144 5.3 Recommendations for Further Study.............................................................. 145

5.3.1 Linear System Theory................................................................................. 145 5.3.2 Training Methodology ................................................................................ 147 5.3.3 Implementation ........................................................................................... 148

BIBLIOGRAPHY........................................................................................................... 150

APPENDIX A: ARMA MODEL .MDL STRUCTURE ................................................ 155

APPENDIX B: STARS IMPLEMENTATION.............................................................. 156

APPENDIX C: FREQUENCY SWEEP PARAMETER SELECTION......................... 160

APPENDIX D: 1D DIVERGENCE DERIVATIONS ................................................... 162

Mach 2.0 ..................................................................................................................... 162 Mach 0.6 ..................................................................................................................... 163

APPENDIX E: STRUCTURAL MODE CONVERSION ............................................. 164

APPENDIX F: SINGLE DEGREE OF FREEDOM DIVERGENCE............................ 165

Configuration Files ..................................................................................................... 165 Modeshape Vector File ............................................................................................... 166

APPENDIX G: PANEL FLUTTER ............................................................................... 167

Plate Parallel Chirp Training Responses at Mach 2.0 ................................................ 167

vii

LIST OF TABLES

Table 3.1 1st Derivative Comparison for Taylor series and ARMA coefficients ......... 92 Table 3.2 0th Order Forcing Function Structural Parameters ........................................ 93 Table 4.1 Single Degree of Freedom Structural Parameters....................................... 105 Table 4.2 Single Degree of Freedom Modal Parameters ............................................ 105 Table 4.3 AGARD 445.6: Experimental Flutter Boundary ........................................ 119

viii

LIST OF FIGURES

Figure 1.1 Aerodynamic Fluid-Structure Interactions Flow Diagram ......................... 2 Figure 1.2 Modeshapes................................................................................................. 5 Figure 1.3 Modeshape Dynamics................................................................................. 5 Figure 1.4 Generalized Forces and Modeshapes.......................................................... 6 Figure 1.5 CFD Flow Chart [Cowan, 2003]................................................................. 7 Figure 1.6 CFD Boundary Conditions: Actual and Transpiration ............................... 7 Figure 1.7 System Identification .................................................................................. 8 Figure 2.1 Wagner Response: Theory and Compressible Results ............................. 11 Figure 2.2 Theodorsen Function ................................................................................ 13 Figure 2.3 Coupled Aeroelastic System..................................................................... 28 Figure 2.4 Aeroservoelastic System Diagram............................................................ 30 Figure 3.1 Classical ARMA Form [Boziac, 1979]..................................................... 39 Figure 3.2 Canonic ARMA Form [Boziac, 1979]...................................................... 39 Figure 3.3 System Identification Flow....................................................................... 45 Figure 3.4 SVD Data Flow......................................................................................... 45 Figure 3.5 CFD, Model and Physical Timescale Relationships................................. 46 Figure 3.6 Discrete Time Root Locus ........................................................................ 48 Figure 3.7 Splice System Identification Flow............................................................ 51 Figure 3.8 Multistep: Motion ..................................................................................... 56 Figure 3.9 Multistep: PSD.......................................................................................... 57 Figure 3.10 Variable Amplitude Multistep: Motion .................................................... 58 Figure 3.11 Chirp: Motion............................................................................................ 60 Figure 3.12 Chirp: Power Spectral Density.................................................................. 61 Figure 3.13 DC-Chirp: Motion..................................................................................... 63 Figure 3.14 DC-Chirp: PSD ......................................................................................... 64 Figure 3.15 DC-Chirp: Motion Determination Example ............................................. 65 Figure 3.16 Fresnel Chirp: Motion............................................................................... 66 Figure 3.17 Fresnel Chirp: PSD ................................................................................... 67 Figure 3.18 Schroeder Sweep: Motion......................................................................... 69 Figure 3.19 Schroeder Sweep: PSD ............................................................................. 70 Figure 3.20 Schroeder Sweep: Peak Factor Comparison ............................................. 71 Figure 3.21 Noise: Displacement ................................................................................. 72 Figure 3.22 Noise: PSD 500 Samples .......................................................................... 72 Figure 3.23 Noise: PSD 5000 Samples ........................................................................ 73 Figure 3.24 Envelope Form.......................................................................................... 75 Figure 3.25 Superposition of Multiple Signals: PSD Holes......................................... 76 Figure 3.26 Noise Experiment: Noisy and Clean Input Signal .................................... 78 Figure 3.27 Noise Experiment: Transfer Function with a Clean Input Signal............. 79 Figure 3.28 Noise Experiment: Transfer Function with a Noisy Input Signal............. 79

ix

Figure 3.29 Noise Experiment: Eigenvalues with Clean Input Signal......................... 80 Figure 3.30 Noise Experiment: Eigenvalues with Noisy Input Signal......................... 80 Figure 3.31 Chi Square: Typical Two Mode Plot ........................................................ 84 Figure 3.32 Force RMS: Two Mode Testcase ............................................................. 85 Figure 3.33 PACF: Aerodynamic System with Chirp Training Signal........................ 86 Figure 3.34 Eigenvalue Explanations........................................................................... 87 Figure 3.35 Model Order Sensitivity Explanation ....................................................... 88 Figure 3.36 Model Order Sensitivity Convergence...................................................... 89 Figure 3.37 Eigenvalue Sensitivity Study Explaination............................................... 89 Figure 3.38 Zero Order Force Training Signal and Output.......................................... 91 Figure 3.39 Output Fit Parameters ............................................................................... 92 Figure 3.40 Zero Order Forces: Eigenvalues ............................................................... 94 Figure 3.41 Zero Order Forces: Zoomed Eigenvalue Crossing ................................... 94 Figure 3.42 Zero Order Forces: System Identification Eigenvalues. ........................... 95 Figure 3.43 2nd Order Force: Continuous-Time Root Locus........................................ 96 Figure 3.44 2nd Order Force: Discrete-Time Root Locus............................................. 97 Figure 3.45 2nd Order Force: Dynamic Input and Output ............................................ 98 Figure 3.46 2nd Order Force: RMS Error Study ........................................................... 99 Figure 3.47 2nd Order Force: Divergence Boundary Study.......................................... 99 Figure 3.48 2nd Order Force: Root Locus Study ........................................................ 100 Figure 3.49 2nd Order Force: Root Locus Study with a Large Model........................ 101 Figure 4.1 SDOF Divergence Geometry.................................................................. 104 Figure 4.2 SDOF Divergence CFD Surface Tetrahedral Grid ................................. 104 Figure 4.3 SDOF: Correct Displacements................................................................ 106 Figure 4.4 SDOF: Incorrect Displacements ............................................................. 106 Figure 4.5 SDOF: Multistep Training Signal Mach 2.0........................................... 107 Figure 4.6 SDOF: Chirp Training Signal Mach 2.0 ................................................. 107 Figure 4.7 SDOF: DC-Chirp Training Signal Mach 2.0 .......................................... 107 Figure 4.8 SDOF: Multistep Eigenvalues ................................................................ 109 Figure 4.9 SDOF: Chirp Eigenvalues....................................................................... 109 Figure 4.10 SDOF: DC-Chirp Eigenvalues................................................................ 109 Figure 4.11 SDOF: Multistep Zoomed Eigenvalues .................................................. 109 Figure 4.12 SDOF: Chirp Zoomed Eigenvalues ........................................................ 109 Figure 4.13 SDOF: DC-Chirp Zoomed Eigenvalues ................................................. 109 Figure 4.14 SDOF: Free Response at 411 psf ............................................................ 110 Figure 4.15 SDOF: Multistep Training Signal Mach 0.6........................................... 112 Figure 4.16 SDOF: Chirp Training Signal Mach 0.6 ................................................. 112 Figure 4.17 SDOF: DC-Chirp Training Signal Mach 0.6 .......................................... 112 Figure 4.18 SDOF: Schroeder Training Signal Mach 0.6.......................................... 112 Figure 4.19 SDOF: Strict Fresnel Mach 0.6............................................................... 112 Figure 4.20 SDOF: State Space Fresnel Mach 0.6..................................................... 112 Figure 4.21 SDOF: Multistep Model Order Sensitivity ............................................. 114 Figure 4.22 SDOF: Chirp Model Order Sensitivity ................................................... 114 Figure 4.23 SDOF: DC-Chirp Model Order Sensitivity ............................................ 114 Figure 4.24 SDOF: Schroeder Model Order Sensitivity ............................................ 114 Figure 4.25 SDOF: Strict Fresnel Model Order Sensitivity....................................... 114

x

Figure 4.26 SDOF: State-Space Fresnel Model Order Sensitivity............................. 114 Figure 4.27 SDOF: Multistep Eigenvalues Mach 0.6 ................................................ 117 Figure 4.28 SDOF: Chirp Eigenvalues Mach 0.6 ...................................................... 117 Figure 4.29 SDOF: DC-Chirp Eigenvalues Mach 0.6................................................ 117 Figure 4.30 SDOF: Schroeder Eigenvalues Mach 0.6 ............................................... 117 Figure 4.31 SDOF: Strict Fresnel Eigenvalues Mach 0.6 .......................................... 117 Figure 4.32 SDOF: State Space Fresnel Eigenvalues Mach 0.6 ................................ 117 Figure 4.33 AGARD: Planform ................................................................................. 119 Figure 4.34 AGARD: Modeshapes ............................................................................ 119 Figure 4.35 AGARD: Multistep Sensitivity............................................................... 121 Figure 4.36 AGARD: Chirp Sensitivity ..................................................................... 121 Figure 4.37 AGARD: DC-Chirp Sensitivity .............................................................. 122 Figure 4.38 AGARD: DC-Chirp Large Amplitude Sensitivity.................................. 123 Figure 4.39 AGARD: DC-Chirp State Space Methodology Sensitivity.................... 124 Figure 4.40 AGARD: Schroeder Sweep Sensitivity .................................................. 124 Figure 4.41 AGARD: Fresnel Sensitivity .................................................................. 125 Figure 4.42 AGARD: Flutter Boundary..................................................................... 126 Figure 4.43 AGARD: Model and Eigenvalue Sensitivity Study Mach 0.499 ........... 127 Figure 4.44 AGARD: Mach Number and Eigenvalue Sensitivity Study................... 128 Figure 4.45 AGARD: Damping Sensitivity at Mach 0.499 and 1.072 ...................... 129 Figure 4.46 AGARD: Damping Sensitivity ............................................................... 129 Figure 4.47 AGARD: Structural Frequency Sensitivity ............................................ 130 Figure 4.48 Panel: Modeshapes and Frequencies ...................................................... 131 Figure 4.49 Panel: CFD Grid ..................................................................................... 131 Figure 4.50 Panel: Serial Training Signal .................................................................. 132 Figure 4.51 Panel: 2-7 Model ARMA Predictions..................................................... 133 Figure 4.52 Panel: Eigenvalues .................................................................................. 133 Figure 4.53 Panel: Parallel Training Signal ............................................................... 134 Figure 4.54 Panel: Free Response at ρ=0.0160 slug·ft-3............................................. 135 Figure 4.55 Panel: Free Response at ρ=0.0163 slug·ft-3............................................. 136 Figure 4.56 Panel: Free Response at ρ=0.0166 slug·ft-3............................................. 136 Figure 4.57 Wing-Flap: Geometry ............................................................................. 137 Figure 4.58 Wing-Flap: CFD Grid ............................................................................. 137 Figure 4.59 Wing-Flap: Mode 1 Training .................................................................. 138 Figure 4.60 Wing-Flap: Mode 2 Training .................................................................. 138 Figure 4.61 Wing-Flap: Ricatti Gains ........................................................................ 140 Figure 4.62 Wing-Flap: Open Loop ........................................................................... 141 Figure 4.63 Wing-Flap: Closed Loop......................................................................... 141 Figure 4.64 Wing-Flap: Closed Loop Chatter Time History ..................................... 142 Figure 4.65 Wing-Flap: Closed Loop Chatter Eigenvalues ....................................... 142

xi

NOMENCLATURE

ACF Autocorrelation Function

AE AeroElastic

AGARD Advisory Group for Aerospace Research and Development

ARMA AutoRegressive Moving Average

ASE AeroServoElastic

b Airfoil Semi Chord

c Airfoil Chord

CASELab Computational ServoElasticity Laboratory

CFD Computational Fluid Dynamics

Cl Lift Coefficient

Cp Coefficient of Pressure

DC Zero Frequency—Direct Current

Kα Torsional Spring Stiffness

M Mach Number

MIMO Multi-input multi-output

NACA National Advisory Committee for Aeronautics

NASA National Aeronautics and Space Administration

na Number of system identification force terms

nb Number of system identification motion terms

nr Number of Structural Modes

xii

PACF Partial Auto Correlation Function

PSD Power Spectral Density

psf Pounds per square foot

psi Pounds per square inch

q Structural States

q∞ Dynamic Pressure

RMS Root Mean Square

ROM Reduced Order Modeling

SDOF Single Degree of Freedom

SISO Single Input Single Output

STARS STructural Analysis RoutineS

SVD Singular Value Decomposition

U Free Stream Velocity

VAMS Variable Amplitude Multistep

α Angle of Attack

ρ Density

ω Angular Frequency

1

CHAPTER 1

1INTRODUCTION

1.1 Background

Nature creates interesting situations resulting from its complexity. Among these

situations, fluid structure interactions are among the most intriguing and dangerous. From

nature’s complexity, a single dangerous and unstable system is created from two well-

behaved systems. Fluid-structure interactions are extraordinarily common: wind in the

trees, waving hair, singing birds, etc. These are interesting but certainly not dangerous.

Human technology has changed these interactions into mysterious, dangerous and feared

occurrences. Howling wind might be intimidating, but losing an aircraft structural

member is disastrous.

The specific type of fluid structure interactions investigated in this thesis is

aeroservoelasticity. Aeroservoelasticity involves the coupling of three areas:

aerodynamics, structural elasticity and servo controls. Aerodynamics concerns the fluid

flow and resulting forces generated from a shape used in aircraft. Structural elasticity

concerns the relationship between the loading and motion of a solid body. Servo-control

concerns changing system characteristics with system information. All three of these

areas interact with each other to cause fluid-structure interactions. The interactions and

influence flows are summarized in Figure 1.1. Each area influences every other area.

2

Structure

Aerodynamics

Controls

Aeroelasticity

Aeroservoelasticity

Input Output

+++

+

Figure 1.1 Aerodynamic Fluid-Structure Interactions Flow Diagram

Within the field of stability and control, there are the two significant questions.

The questions are: Will the structure break? , and How can I make the structure do what I

want?. These questions reduce to single fundamental question: How can I predict the

system’s response?. The answer contains multiple alternatives: test the actual system,

make a heuristic guess or develop a system model. Developing a system model is often

the most robust and insightful approach.

Developing a system model for aeroservoelastic prediction requires models for

the aerodynamic, structural and servo-controls sub-systems. The structural response

comes from a free-vibration solution with an added forcing function. While the

aerodynamic system model could come from any number of methods, this thesis will use

a computational fluid dynamics (CFD) solver. From the input boundary conditions,

output pressures are determined. The servo-controls sub-system is determined from a

typical controls analysis.

The digital computer established computational fluid dynamics (CFD) as a

powerful engineering tool. However, CFD has disadvantages in determining fluid-

structure interactions. First, CFD only solves one particular set of boundary conditions.

3

Second, CFD requires tremendous amounts of processor power. A computer might spend

a month or more solving one coupled fluid-structure problem. These drawbacks reduce

the effectiveness of CFD as a controls or stability prediction tool.

System identification offers a practical solution to these drawbacks by using CFD

generated data to create an aerodynamics system model. Ideally, this aerodynamics

model will exactly reproduce the aerodynamic outputs for the boundary condition inputs.

By coupling the aerodynamic and structural states, practical analysis of stability and

control is possible.

The research contained in this thesis is conducted for the Computational-Aero-

Servo-Elasticity-Laboratory (CASELab) in the Mechanical and Aerospace Engineering

department at Oklahoma State University. The CASELab supports the STARS group in

flight test operations at NASA Dryden. The STARS group develops and maintains a

group of analysis programs named STructural Analysis RoutineS [Gupta, 2001].

1.2 Current Status

The basic tools required for aeroservoelastic predictions already exist. STARS

contains both structural response and aerodynamics solving routines. The STARS Solids

routine performs the structural response calculations and determines the frequencies and

modal properties for free structural vibration. Aerodynamic forces are computed with a

finite-element, unsteady, Eulerian based fluid dynamics solver. STARS contains two

separate aerodynamics solvers: the old STARS mg2 and the new euler3d. The euler3d

code was developed from Tim Cowan’s dissertation work [2003]. STARS currently

contains a system identification prediction routine for both the old STARS mg code and

4

the euler3d code. The euler3d system identification routine was ported from the old

STARS code and was not validated.

1.3 Simulation Overview

This section describes the fundamentals of the aeroservoelastic prediction

methodology used in this thesis. The underlying process combines structural and

aerodynamic simulations. These two processes are combined for aeroservoelastic

predictions. The final fundamental process is the system identification or training

methodology.

1.3.1 Structure

The structural simulation determines the time response of the elastic structure. For

typical aeroservoelastic behavior, the structure remains in the linearly elastic region. This

allows for decomposing arbitrary motions into a sum of orthogonal modeshapes, Φ, with

individual modal motions, q. Practically, this means that each structural mode has an

associated shape, mass, stiffness and damping. The second order differential equation for

structural motion in the modeshape reference frame is:

[ ] [ ] [ ] )(tFqKqCqM =++ ΦΦΦ &&&

This allows arbitrary structural motions of complex aerospace vehicles to be simulated by

simple second order ordinary differential equations. The structural response is described

by an array of generalized modeshape displacements rather than describing each

individual point on the structure. For example, the first 4 bending modes of a cantilever

beam are plotted in Figure 1.2. Likewise, the dynamic displacement response of a mode

5

is described by a single number, generalize motion. This is shown for first mode bending

in Figure 1.3.

ModeshapeFrequency

3.516

22.03

61.70

120.9

Figure 1.2 Modeshapes

Phase [deg]

180

90

0

150

120

60

30

GeneralizedMotion

1.00

-1.00

0

0.33

0.66

-0.33

-0.66

Structural Displacement

Figure 1.3 Modeshape Dynamics

A complication arises when external forcing is applied. The overall external

forcing must be reduced to a single generalized force. When an arbitrary pressure

distribution is applied, the effective generalized force on each mode is determined by:

),()()( txPxtF TΦ=

This is essentially the dot produce of the local external force and the modeshape. The

concept is illustrated in Figure 1.4 for a spatially constant force. Thus, the effective force

on a mode is reduced to a single number, generalized force.

6

Modeshape and Applied Force

Generalized Force

2.16

0.53

)()( xPxF ⋅Φ=

0.61

0.84

Figure 1.4 Generalized Forces and Modeshapes

The overall governing equation for the structural motion is the following second order,

constant coefficient differential equation:

[ ] [ ] [ ] ),()( txPxqKqCqM TΦ=++ ΦΦΦ &&&

1.3.2 Aerodynamics

The aerodynamic simulation determines the temporal and spatial characteristics of

a fluid flow. The objective is to determine the local pressures within a specified flow field

subjected to specified boundary conditions. For the CFD analysis presented in this thesis,

the governing fluid dynamics equation is the compressible Euler equation. Unsteady flow

solutions are obtained by a time-marched approach. This requires local iterations to

converge the flow solution at each timestep. Global steps advance the solution in time. A

flow diagram of the CFD solver reproduced from the euler3d development dissertation

[Cowan, 2003] is shown in Figure 1.5.

7

read solver control parameters

read geometry data: COOR, IELM, ISEG, IBEL

compute additional geometry data: G2D/G3D, DM, RBE, WSG, ANOR

read any elastic/dynamic data

set/read initial conditions for UN for t = 0

compute initial aerodynamic loads for t = 0

compute initial structural dynamics state for t = 0

output initial conditions for t = 0

UNO = UN

UN1 = UN

do istp = 1,nstp

advance structural dynamics from t = n to t = n + 1

update ANOR and compute BVEL (transpiration) for t = n + 1

compute local time step, DELT, for each node

do icyc = 1, ncyc

initialize RHS

enforce flow tangency on UN1

add element integrals to RHS

add boundary integrals to RHS

add dissipation to RHS

enforce flow tangency on RHS

UN1 = UN1 – DELT·DM·RHS

enddo

output solution residuals

UN0 = UN

UN = UN1

compute new aerodynamic loads for t = n + 1

output forces and dynamics for t = n + 1

if MOD(istp, nout) = 0, output solution unknowns

enddo

Figure 1.5 CFD Flow Chart [Cowan, 2003]

Unsteady structural boundary conditions are specified from the structural modeshapes, Φ,

using transpiration. Transpiration simulates an actual structural motion with a change in

the boundary normal. The actual CFD grid does not move. Transpiration is illustrated for

a displacement boundary condition in Figure 1.6.

SteadyBoundaryCondition

UnsteadyBoundaryCondition

Actual Boundary Conditions Simulated Boundary Conditionswith Transpiration

Figure 1.6 CFD Boundary Conditions: Actual and Transpiration

8

Integration of the local pressure over the structure allows for calculating the generalized

forces. Aeroservoelastic responses are determined though a loop. The CFD solver

determines generalized forces given generalized motions. The structural solver

determines new generalized motions given the generalized forces. The entire process is

then repeated. This process is stylized in Figure 1.1.

1.3.3 System Identification

The objective of system identification is to determine a system model that predicts

the system dynamics. For this thesis, system identification involves determining

generalized forces from generalized motions. The process is conceptually shown in

Figure 1.7.

Input Known Motion

Excite UnsteadyAerodynamics

Create SystemModel

GeneralizedMotion

GeneralizedForce

System ModelCoefficients

[ ]=ia

System Flow Diagram

Typical Output

Output Data Type

Figure 1.7 System Identification

The basic process is to input a known motion into the aerodynamic system. The CFD

solver uses this input generalized motion to calculate generalized forces. The relationship

between the generalized motions and forces is used to create a system model. Once the

system model is determined, the generalized forces resulting from arbitrary generalized

motion inputs can be found.

9

1.4 Objectives

The objective of this thesis is to improve the current state of linear system

identification techniques for aeroservoelastic predictions. The system under consideration

is multi-input multi-output aerodynamics for use in stability and control analysis.

Limitations and sensitivities with the present prediction method will be identified.

These critical areas will be found by evaluating the current prediction method with

known physics and computational limitations.

Development of improved prediction techniques will follow from the experience

gained. Prediction selection criteria will be developed to evaluate model performance.

These improvements will be implemented and tested.

Finally, the improved prediction methodology will be used for stability and

control analysis of typical aeroservoelastic systems. Results will be compared with the

old methodology. Questions and further research suggestions will be presented.

10

CHAPTER 2

2LITERATURE REVIEW

This chapter examines the literature for unsteady aerodynamics and description

techniques. Some fundamental unsteady aerodynamic theories are reviewed. Then, a

review of relevant system identification literature is given. Next, excitation signal

literature is reviewed. Finally, previous work concerning the development of the

structural, aerodynamic and aeroservoelastic system models is reviewed.

2.1 Unsteady Aerodynamics

A fundamental requirement of aeroservoelastic modeling is determining the

forces caused by unsteady aerodynamics. The boundary condition history specifies the

flow geometry; however, a complication arises because the fluid flow has memory. This

makes determining the forces resulting from unsteady fluid flow significantly more

complex.

The general problem for unsteady aerodynamics is to find a solution that is

consistent with fluid continuity, fluid momentum conservation and the unsteady

boundary conditions. In general, a closed-form solution to arbitrary unsteady flows is not

available.

Closed form simplified solutions to supersonic unsteady flows are available.

Supersonic flows are tremendously simplified because flow influences cannot travel

upstream [Dowell, 1995]. For small perturbations, piston theory provides a simple and

11

reasonably accurate flow prediction methodology [Hunter, 1997]. This solution

methodology is already implemented in STARS.

Closed form subsonic unsteady solutions based on simplified physics are

available when assuming certain conditions. The typical assumptions are inviscid,

irrotational and incompressible flow with a thin airfoil. These assumptions reduce the

solution to an analytic solution, which can be solved for certain restrictive boundary

conditions. The classic Wagner and Theodorsen unsteady solutions result from these

assumptions.

The unsteady Wagner problem consists of an impulsively started airfoil operating

in an incompressible inviscid flow. At time zero, the airfoil undergoes a step change in

angle of attack. The governing equations are Kelvin’s theorem and Laplace’s equation

with the properly defined boundary conditions. The time response for the Wagner

problem is shown in Figure 2.1. For this incompressible flow, “any change of boundary

conditions is propagated instantaneously...” [Bisplinghoff, 1955].

0 1 2 3 4 5 6 7 8 9 100

0.2

0.4

0.6

0.8

1

Chords Traveled

Uns

tead

y Li

ft R

atio

Mach 0.3

Mach 0.8

Steady State

Wagner

Figure 2.1 Wagner Response: Theory and Compressible Results

12

The Wagner problem is not a physically realizable compressible aerodynamic

operation due to the step change in the boundary conditions. A CFD solver can solve the

flow field for this step input; however, the flow solution is not actually the desired step

input response. Parameswaren [1997] describes this situation:

For example, if the airfoil is suddenly exposed to the flow at a different

angle of attack, then the airfoil will also experience an infinite pitch rate.

Then, the response will certainly be a reflection of a step change in

angle of attack as well as an impulsive change in the pitch rate.

Experimentation with the STARS flow solver shows that a “step” angle of attack input

also excites pitch rate and acceleration terms. Figure 2.1 shows the corresponding Mach

0.3 and Mach 0.8 compressible Wagner responses as computed by euler3d. These results

qualitatively agree with the compressible indicial lift response calculated by Bisplinghoff

[1955]. Infinite lift is predicted for the 0+ timestep. This indicates that the Wagner step

input will not properly excite the desired aerodynamic response.

The Theodorsen problem describes the unsteady aerodynamics resulting from an

airfoil undergoing harmonic pitching and plunging motion. The classical derivation is

made in an incompressible inviscid flow [Bisplinghoff, 1955].

The expression for the magnitude of the harmonic lift is:

( ) ( ) )(22 kCbaUhUbbaUhbL ααρπααπρ &&&&&&& −++−+=

The C(k) term is a frequency dependent wake influence term. The parameter k is defined

as a reduced frequency in the following form:

Uck

2ω=

13

The Theodorsen function, C(k), is complex valued and is defined as the following ratio of

Hankel functions:

)()()(

)( )2(0

)2(1

)2(1

kiHkHkHkC

+=

A plot of the Theodorsen function is given in Figure 2.2. For low frequencies, this

form approaches the steady state lift. At high frequencies, the unsteady lift magnitude

approaches one half the steady state lift.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0.2

0.1

0

0.1

0.10.20.3

0.5

1

2

4

∞ 0

Theodorsen Function, C(k)

Reduced Frequency

U

ck

2

ω=

Real C(k)

Imaginary C(k)

Figure 2.2 Theodorsen Function

It is important to notice that the maximum unsteady phase difference occurs near

a reduced frequency of 0.2. At high and low reduced frequencies, the unsteady

aerodynamic loading is in phase with the motion. This implies that for an airfoil-based

aerodynamic system, the largest phase differences occur at moderately low reduced

frequencies. Destructive aero-structural coupling typically occurs in the moderate to low

reduced frequency range.

The unsteady aerodynamics resulting from arbitrary motion can be described by

combining Fourier theory and the Theodorsen result. The relationship between the time

domain and frequency domain representations of a signal is based on Fourier theory.

Given a generalized motion function f, the lift in the time domain is [Zwillinger, 1996]:

14

∫∞

∞−

⋅−= ωρπρ ω dekfkCUbtfbL ti)()()(2

This implies that a general aerodynamics function will have force components resulting

from both instantaneous motion and lagged wake convection.

2.2 Aerodynamic System Models

This section compares description functions for aerodynamic modeling. There are

many possibilities to choose from. A good overview of the possibilities is given by Ljung

[1987]. Several system models are compared: ARMA, indicial, reduced order modeling

and nonlinear. The basic comparison will be between the ARMA and indicial methods.

Nonlinear methods are only discussed as a comparison to linear methods.

2.2.1 Indicial Methods

Indicial models describe outputs by the convolution of responses and inputs.

These models have output expressions derived from the following continuous time

convolution integral [Chen 1999]:

∫ −=t

dtutgty0

)()()( ττ

The indicial response function is g(t) and the input is u(t). For systems with a single input

parameter, the indicial response is simply the step response. For coupled input systems,

the indicial response becomes significantly more complicated. For a proper indicial

response, the effects of displacement, velocity and the higher order terms must be

decoupled. This can be difficult in aerodynamic training. Additionally, this requires an

indicial response for every proposed input term. For example, Bisplinghoff [1955] shows

that the compressible indicial lift response starts from infinity, which is in contrast to the

15

incompressible Wagner response. The difference is the presence of coupled boundary

conditions. Indicial response functions must be specifically decoupled and trained for

each motion boundary condition. This is a major disadvantage for generic aerodynamic

predictions.

Aerodynamic representation by indicial methods is popular. Cai [2000] used

indicial aerodynamic responses on displacements and velocities for bridge aeroelastic

response calculations. The indicial responses were calculated by fitting to Theodorsen-

based frequency-domain expressions. Cai admits that the indicial response is “very

difficult to measure directly.” Reisenthel [1997] extended the basic indicial response to

account for nonlinear effects in aircraft flight dynamics.

2.2.2 ARMA

The ARMA form consists of a discrete time input-output linear relationship

[Brockwell, 1991]. The basic premise is that the output depends on that previous outputs

and inputs. At the kth timestep, the mathematical representation of the output is the

following:

[ ] [ ]∑∑−

==

−+−=1

01)()()(

nb

ii

na

ii ikxBikyAky

The system’s input response is described by:

[ ]∑−

=

−1

0)(

nb

ii ikxB

The system’s internal response is described by:

[ ]∑=

−na

ii ikyA

1)(

16

The ARMA form directly corresponds to a discrete form of an ODE. The general

form for an ARMA model can be written as a p-order output and a q-order input.

∑∑ −=− )()( qtxbptya qp

A general ODE is written as an m-order output and an n-order input.

( ) ( )∑∑ = nn

mm xdyc

Expanding the ODE expression for a 4 by 3 order yields the following result.

( ) ( ) ( ) ( ) ( ) ( ) ( )22

11

00

33

22

11

00 ycycycycycycyc ++=+++

Using a Taylor’s series approximation to derivatives yields the following finite difference

expressions.

( )

( )

( )t

kyy

tkykyy

kyy

∆+≈

∆−−≈

=

L)(

)1()()(

2

1

0

( )

( )

( )t

kxx

tkxkxx

kxx

∆+≈

∆−−≈

=

L)(

)1()()(

2

1

0

Substituting in these finite difference approximations yields the following expression.

LLLLLL +−

+

∆−

+

+

∆+=+−

+

∆−

+

+

∆+ )1()()1()( 11

011

0 kxtdkx

tddky

tcky

tcc

This is exactly the ARMA form.

The ARMA form assumes a linear relationship between the current output and the

previous states. This form allows for simplified system identification. The following

expression shows the overall linear expression:

}OutputCurrent InputsCurrent andPast

11

Outputs Previous

)()1()1()()()()1(M

44444444444 844444444444 76

MLM

LLL4444 84444 76

M

L tyBAnbtxnbtxtxtxnatyty

i

inrnr =

−−−−−−

In the general case, this expression reduces to the following linear algebra form:

17

[ ] bxArr =

This form can be solved through any of a number of linear algebra methodologies.

The advantages of ARMA are tremendous. First, ARMA allows for an intuitive

system identification implementation. The model coefficients can directly be determined

from linear system theory. Second, the ARMA form is easily transformed to a state space

form consistent with aerodynamic prediction requirements. Third, ARMA can properly

model MIMO systems without additional complications.

A particular disadvantage of ARMA is determining a sufficient model order. This

model order problem causes particular difficulties for long lag systems. Multiple papers

give model order estimation guidelines; however, most are based on difficult to

implement stochastic frameworks. Gevers [1986] derives the McMillan degree, order, of

an ARMA system. Gevers states:

These [monic ARMA] models can only represent systems whose

McMillan degree is a multiple of the number of outputs. In all other

cases they will tend to produce estimated models of higher order than

the true system.

This result has not been discussed in the literature with respect to aerodynamic modeling.

2.2.3 Nonlinear

Nonlinear description models allow for advanced predictions for complex

systems. The number of methods, which are available for nonlinear descriptions, is

immense. These methods allow for advanced predictions by capturing more of the

underlying physics. For aerodynamics, nonlinear analysis is needed for limit cycle

predictions as well as high angle of attack and other common flow regimes. Hamel

18

[1996] gives examples of nonlinear unsteady flow models and their performance results.

Theoretically, a nonlinear description should have no small-perturbation limitations. In

practice, assuming a nonlinear description complicates the system identification process.

Nonlinear analysis is already computationally expensive. Additionally, the nonlinear

identification process requires data throughout the desired system prediction operational

area [Reisenthel, 1997]. If the training does not excite the dominant nonlinearities, then

the identification will not determine an accurate system model. This may require a

training length an order of magnitude longer. A useful excitation will need information at

all expected motion magnitudes. Depending on the system complexity, the excitation

may need exotic motion specifications to properly account for the nonlinear system

dynamics. Some of these motion combinations are difficult to reach in a consistent

manner. Another disadvantage is the loss of traditional linear system theory. This implies

that the entire identification and evaluation process becomes time response dominated.

All response evaluations will be based on how the system reacts in time. This forces

system identification into more of a black-box experiment rather than a system property

evaluation. Prematurely switching to nonlinear models appears to inhibit the

understanding and physical insights allowed by linear system identification. For many

aerostructural problems, the fundamentals can be captured by a linear system. As Dowell

[1996] points out:

One of the pleasant ironies for the aeroelastician has been that such

calculations have revealed that a linear dynamic model perturbed from a

nonlinear static or steady flow model which includes shock waves is

sufficiently accurate for describing many flutter phenomena.

19

A related nonlinear description technique is reduced-order-modeling. This

technique seeks to reconstruct the unsteady flow properties through an eigenvalue

analysis of the entire flow field. Not surprisingly, ROM is tremendously computationally

expensive. Also, ROM does not result in a direct state space form for aerodynamic

forces; the resulting pressure must be integrated over the modeshape. ROM does not

appear to be a general method for aerodynamic predictions due to jump changes in

aerodynamic flow solutions [Raveh, 2001] [Reisenthel, 1997]. Florea [2000] found that

ROM gave excellent flow predictions even at transonic Mach numbers. Florea found that

a 2D NACA 0012 airfoil at Mach 0.1 required 10 eigenvalues while the Mach 0.85 case

required approximately 310 eigenvalues, which presented further computational

difficulties.

2.3 System Identification Methodologies

The objective of system identification is to fit a system model to measured data.

Better system models result from more data; however, this effectively creates an

overdetermined system, which has more data than equations. Björck [1996] states, “... the

solution which minimizes a weighted sum of the squares of the residual is optimal in a

certain sense.” The system model as presented in the previous section can be reduced to a

general linear equation form as shown below:

[ ] bxArr =

Direct solutions by the traditional linear algebra techniques will not be successful

for multiple reasons. First, [ ]A is not square; a traditional inverse does not exist. Second,

the condition of [ ]A is typically such that the numerical routine will create significant

errors. A better solution technique is to switch to a methodology capable of handling

20

these difficulties. Iteration and Singular Value Decomposition appear to be the most

commonly used numerical routines for linear system identification. The current STARS

system identification method uses SVD.

Iteration consists of solving a particular equation with successive approximations.

The routine consists of iteration until the solution is “good” enough. Björck shows that

iteration is “attractive” for sparse matrices due to storage requirements, convergence rate

and solution speed. Improved solution speed typically results from a compressed matrix

form [Björck, 1996]. Multiple iterative solution methods exist; picking one method

involves selecting among many schemes, pre-conditioners and factorizations. Iteration

also allows for advanced solutions based on external constraints and other non-typical

least squares formulations. The primary disadvantage of iteration schemes is their

consistency. Björck [1996] notes:

The main weakness of iterative methods is their poor robustness and

often narrow range of applicability. Often, a particular iterative solver

may be found to be very efficient for a specific class of problems, but if

used for other cases it may be excessively slow or break down.

This is a significant disadvantage for system identification.

Singular value decomposition is a linear algebra technique for reducing any

matrix to two basis matrices and one diagonal matrix. The SVD expression form for an

arbitrary matrix A is given as [Press, 1992]:

SVUA T=

The U and V matrices are the basis matrices; the diagonal S matrix contains the singular

values. The pseudoinverse of A is [Press, 1992]:

21

TVWUA =+

W is the inverse of each diagonal term in the S matrix. Solution simplifications are

possible by dropping small singular values. Press [1992] proposed using this

simplification to determine the dominant solution structure.

The advantages of SVD are significant. SVD always finds a least-square solution.

Press says, “SVD can be significantly slower than solving the normal equations;

however, its great advantage, that it cannot fail, more than makes up for the speed

disadvantage”. Another advantage is that the SVD data structure shows solution

information based on the singular values. Björck [1996] shows that “perturbations of the

elements of a matrix produce perturbations of the same, or smaller, magnitude in the

singular values.” The primary disadvantage of SVD is the computational expense,

especially for large problems. Also, small singular values are often corrupted by machine

precision errors.

Variants of the SVD routine exist for specialized applications. Traditional SVD

provides a “most” least-square solution [Press, 1992] based on the Frobenius norm, L2.

The Frobenius norm is defined as:

∑= 2

2 iaA

Others have demonstrated that the traditional least-squares formulation does not

always yield the best solution. A simple modification to the traditional SVD routine is the

Total Least Squares. Deprettere [1988] shows that the method “is appropriate when there

are errors in both the observation vector b and the data matrix a...” Björck [1996] states

that Van Huffel [1991] found 10 to 15 percent improvements in solution quality with

Total Least Squares over Least Squares. The solution quality criteria were not reported.

22

Björck [1996] also show a methodology for converting Total Least Squares problems to

Least Squares problems.

Subspace identification techniques are popular. Moonen [1990] proposed a

“quotient SVD” for system identification with non-white noise. The solution

methodology reduced to performing SVD twice. The additional SVD calculation

determines an input-output data subset with the greatest signal to noise ratio. The results

show that QSVD gave improved predictions for small signal to noise ratios. Overschee

[1993] uses a similar method to compute a system model based on a Kalman filter, which

was based on the raw input-output data. The approach is significantly more complicated

than the traditional SVD approach. Additionally, the approach appeared to give more

precise solution but with a significantly wider error distribution when compared to a

traditional approach.

For computer vision applications, Irani [2002] found that including uncertainty

into the SVD formulation significantly improved shape and motion predictions.

Directional uncertainty estimates were included through a covariance weighted scheme,

which minimized the Mahalanobis distance rather than the Frobenius norm. Mahalanobis

distance effectively assigns variable penalty weighting factors to each orthogonal solution

directions. This means that errors in the x direction, for example, are penalized more than

errors in the y direction. The disadvantage of this formulation is that useful penalty

weighting factors are needed. Determining the weighting factors can become complicated

for multi-dimensional problems. For Irani, a useful covariance data structure resulted

from the physical restriction of four dimensions: space and time.

23

2.4 Input Training Signals

Successful system identification requires selecting a sufficiently good input

training signal. Selection techniques are typically based on estimated system

characteristics. Typically, these criteria are only mathematical in form and are not tested

for practicality and sensitivity. This section will review basic input signal forms and

examine their performance for aerodynamic system identification in the time domain.

The literature provides abundant information on signal design and system

identification techniques. However, most input signals are designed and evaluated for

frequency domain identification. While the fundamental requirements for the time

domain are similar, identification in the time domain has some subtle differences. Any

input signal can be used for time domain identification [Miller, 1986]; however, the

quality of the resulting model will reflect the input signal quality. While basic system

predictions may be possible with simple system models and arbitrary input signals,

system models that are used for controls need good resolution and accuracy in describing

the system. The input signal needs to take into account the overall system processes to

avoid over-driving or forcing non-linearity [Braun, 2002]. Historically, the binary signal,

the frequency sweep and the multisine are the most commonly used signals for system

identification. For flight-testing, the multisine and the binary signals are typically the

most common [McCormack, 1995]. The binary signal, the frequency sweep and the

multisine will be investigated.

Binary signals consist of a series of pulses arranged for a specified frequency

spectrum. Unfortunately, the binary signal frequency spectrum contains unexcited areas

that can only be removed by adding significant length to the signal [Schoukens, 1988]. In

24

this regard, binary signals are similar to noise or stochastic signals [Schoukens, 1988].

Random noise signals could be considered as binary signals with variable pulse heights.

Schoukens [1988] shows that random noise does not automatically average out system

nonlinearities. A binary signal, the 3211 multistep, is one of the most commonly used

flight-test signals [Miller, 1986]. This may be in part because the multistep is easy to

implement and contains a relatively simple functional form.

Frequency sweeps are generated from sinusoidal functions with a smoothly

varying frequency. This has the advantage of exciting all frequencies within a specified

bandwidth. Linear frequency sweeps are commonly referred to as chirps. Frequency

sweep are common in the flight test community [Miller, 1986]. However, one

disadvantage of a sweep is poor low-frequency performance [Young, 1990][Brenner,

1997]. Investigations into improving chirp frequency response are not common; most

investigators switch to multisines to improve low frequency response. Numerical

difficulties [Zwillinger, 1996] associated with advanced frequency sweep forms are,

possibly, responsible for limited development. Another disadvantage is that the frequency

sweep can overexcite the structure and cause ``critical flight incidences’’ [Miller, 1986].

Sweeping through a structural or aeroelastic resonance frequency allows for the

possibility of unintentionally overexciting the structure. Because the sweep advances in

frequency, the structure may already have dangerous motion amplitudes even before the

actual resonance frequency.

The classic multisine signal is the Schroeder sweep [Schroeder, 1970]. Variable-

phase discrete frequencies are added to yield an arbitrary PSD across a specified

bandwidth. The Schroeder sweep visually resembles a frequency sweep during part of the

25

overall harmonic signal [Van Der Ouderaa, 1988]. Algorithms for minimizing the

multisine peak factor are common [Braun, 2002] [Van Der Ouderaa, 1988] [Mehra, Dec.

1974] [Mehra, June 1974]. For overall performance, Young and Patton [1990] found that

for a specific helicopter identification problem, a multisine gave slightly better results

than the corresponding frequency sweep. This improvement appeared to be due to

improved low frequency excitation with the multisine. Simon [2000] found that the

Schroeder sweep is sensitive to the signal excitation length.

2.5 Model Quality

Model quality evaluations are necessary to ensure the system identification

process correctly modeled the relevant physics. An excellent criterion exists for

statistically based models [Akaike, 1974]; however, it is not quite as intuitive for ARMA

based models. Ninness [1995] surveys model quality estimation methodologies for both

frequency domain and time domain identification. The underlying emphasis is model

selection and identification for robust control. Ninness indicates that deterministic

identification is preferred over stochastic identification for robust control. Mäkilä [2003]

provides a methodology for determining the size of unmodeled dynamics between a

linear system and the actual system. The paper proposes using the Fréchet derivative to

evaluate mildly nonlinear systems. Model quality is affected by the discrete time

timestep. “It is well known” [Worden, 1995] that over-sampling leads to system

identification difficulties. Åström [1969] shows that an optimal sampling rate occurs for

simple system identification. As expected, Åström found that the model quality

degradation is worse for undersampling than for oversampling.

26

2.6 Aerodynamic and Structural System Representations

The aeroelastic predictions are based on aerodynamic and structural models. This

section will discuss the formulation of these models. These derivations were created by

Cowan [2001].

2.6.1 Structural Model

Structural motion is decomposed into orthogonal modes. The overall structural

response becomes a second order ordinary differential equation with constant structural

coefficients M, C and K and an arbitrary modal forcing function F:

[ ] [ ] [ ] FqKqCqM =++ &&&

Two structural states are required to represent this system. Each state contains the

displacement and velocity of each mode. Mathematically, the structural state vector is

given as the following:

[ ]Ts qqx &r =

Decomposing the second order differential equation into two first order equations yields

the following state space form:

[ ] [ ][ ] [ ] )()()(

)()()(tFDtxCtqtFBtxAtx

sss

ssss

+=+=&

The state space parameters As, Bs, Cs and Ds are determined from the structural

parameters by the following matrix form:

[ ] [ ] [ ][ ] [ ] [ ] [ ]

−−

= −− CMKMI

As 11

0 [ ] [ ]

[ ]

= −1

0M

Bs

[ ] [ ] [ ][ ]0ICs = [ ] [ ]0=sD

27

The above continuous state space form is transformed to a discrete state space

form with the matrix exponential. This form is based on an evenly spaced discrete

timestep. The state space form becomes the following at the kth time step:

[ ] [ ][ ] [ ] )()()(

)()()1(kFDkxCkq

kFHkxGkx

sss

ssss

+=+=+

The state space parameters Gs, Hs, Cs and Ds result from the appropriate discrete

transformations:

[ ][ ] [ ] [ ]( )[ ] [ ]ssss

dtAs

BAIGH

eG1−

−=

=

The structural response is now represented in a discrete state space form appropriate for

connecting as an input-output system.

2.6.2 Aerodynamic Model

The aerodynamics are represented as a linear ARMA state space model.

Constructing the aerodynamic system model requires storing previous aerodynamic

forces f and inputs q . The system model state vector is the following:

+−

−−

=

)1(

)1()(

)1(

)(

nbkq

kqnakf

kf

kxa

M

M

In a discrete state space form, the aerodynamic state space form is the following. This

aerodynamic model was composed in discrete time, so a continuous-discrete conversion

is not necessary as it was in the structural case:

28

[ ] [ ][ ] [ ] 0)()()(

)()()1(fkqDkxCkf

kqHkxGkx

aaaT

aaa

++=+=+

The following state space parameters Ga, Ha, Ca and Da result from expanding the ARMA

state space model in terms of the state space vector )(kxa :

[ ]

[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ]

[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ]

[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ]

=

−−−

0000000

00000000000000000000000000000

00000000000000

1221121

I

II

I

II

BBBBAAAA

G

nbnbnana

a

LL

MMOMMMMOMM

LL

LL

LL

LL

MMOMMMMOMM

LL

LL

LL

[ ]

[ ][ ][ ]

[ ][ ][ ][ ]

[ ]

=

0

00

0

00

0

M

M

I

B

H a

[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ]1221121 −−−= nbnbnanaa BBBBAAAAC LL [ ] [ ]0BDa =

2.6.3 Aeroelastic Model

The above development of discrete time structural and aerodynamic system

models is continued by joining the models. The coupled system diagram is given in

Figure 2.3.

Structure

Aerof̂

q

∞q

f

TfIf+

+

Figure 2.3 Coupled Aeroelastic System

The resulting state space model describes the system’s coupled aeroelastic response. The

overall structural response vector is given by )(kq . The aerodynamic forces are scaled by

29

the dynamic pressure ∞q . The system model contains an impulsive forcing function via

the If input.

Substitution yields the following system model:

[ ][ ][ ]

=

+

+

+=

++

∞∞∞

)()(

0)(

0

ˆ

0)()(

)1()1( 0

kxkx

Ckq

fHqfH

kxkx

GCHCHqCDHqG

kxkx

a

ss

sI

s

a

s

asa

assass

a

s

This coupled form of the combined aerodynamic and structural models is

sufficient for determining aeroelastic free response characteristics. A free response

simulation consists of matrix multiplication and state vector storage. Matrix operations

are orders of magnitude faster than an unsteady CFD solution. The system stability

boundary can be found in the time domain by successive bounding of stable-unstable

dynamic pressures.

The intrinsic system stability is determined from the plant matrix. The plant

matrix is a function of the structure with the submatrices Gs, Hs and Cs and a function of

the aerodynamics with the submatrices Da, Ca, Ha and q∞. The plant matrix shown below

has dimensions ( )( ) ( )( )11 ++⋅×++⋅ nanbnrnanbnr :

+ ∞∞

asa

assass

GCHCHqCDHqG

Locations of eigenvalues in the z plane with respect to the unit circle indicate the overall

stability of the system. There are ( )1++⋅ nanbnr eigenvalues for a na-nb model with nr

modes.

30

2.6.4 Aeroservoelasticity Model

The aerodynamics and structural system discrete-time system models allow for

the addition of a control system with pilot inputs. A typical system diagram for this

aeroservoelastic system is given in Figure 2.4.

StructureAerof̂ q

∞qf Tf

If

++

q

eδq +

+

Control

cq+

Iq

Figure 2.4 Aeroservoelastic System Diagram

Multiple control methods are possible. For output feedback, a simple control law

is determined by choosing a control gain matrix K:

[ ] )()( kqKkqc =

Derivation of the resulting state-space form requires coupling four sub-systems:

structures, aerodynamics, controls and aeroservoelastic coupling. The structures and

aerodynamics systems in state-space form are repeated from sections 2.6.1 and 2.6.2. The

aerodynamics state-space representation is:

[ ] [ ][ ] )()(

)()()1(kxCkq

kfHkxGkx

ss

Tssss

=+=+

The structural state-space representation is:

[ ] [ ][ ] [ ] 0

ˆ)()()(ˆ)()()1(

fkqDkxCkf

kqHkxGkx

aaa

aaaa

++=

+=+

The aeroservoelastic coupling terms result from joining the structural and control

systems:

31

qqeqqq

fqf

fff

cI

IT

+=+=

=

+=

δ

δ

ˆ

The derivation of the aeroservoelastic state-space form is as follows. The aeroservoelastic

coupling terms are added into the aerodynamics:

0ˆ)()()(ˆ

)()()1(

fkeDkxCkf

keHkxGkx

aaa

aaaa

++=

+=+

Next, the structural outputs are added to the resultant aerodynamics:

( )( ) 0

ˆ)()()()()(ˆ)()()()()1(

fkxCkqkqDkxCkf

kxCkqkqHkxGkx

sscIaaa

sscIaaaa

++++=

+++=+

The control algorithm is added to the aerodynamics:

( )( ) 0

ˆ)()()()()(ˆ)()()()()1(

fkxCkxKCkqDkxCkf

kxCkxKCkqHkxGkx

ssssIaaa

ssssIaaaa

++++=

+++=+

Next, like terms are collected:

[ ][ ] 0

ˆ)()(I)()(ˆ)()(I)()1(

fkqDkxCKDkxCkf

kqHkxCKHkxGkx

Iassaaa

Iassaaaa

++++=

+++=+

Now, the coupling terms are added to the structural state-space form:

( ))()(ˆ)()1( kfkfqHkxGkx Issss ++=+ ∞

The aerodynamic system is combined with the structural system, and like terms are

collected:

[ ]( )Iss

Iasaasssasss

fHfHq

kqDHqkxCHqkxCKDHqGkx

++

++++=+

∞∞∞

)()()(I)1(

Finally, the resultant aerodynamics and structural forms are combined into a single state-

space form:

32

[ ][ ]

[ ][ ][ ]

=

+

+

+

+

++=

++ ∞∞∞∞

)()(

0)(

ˆ00)(

)(I

I)1()1(

0

kxkx

Ckq

fHq

qH

DHqf

Hkxkx

GCKHCHqCKDHqG

kxkx

a

ss

sI

a

asI

s

a

s

asa

assass

a

s

This form is sufficient for aeroservoelastic predictions in the time domain. Again, the

plant matrix can be evaluated for eigenvalue placement.

33

CHAPTER 3:

3METHODOLOGY

In this chapter, techniques for improved aeroservoelastic predictions are

developed and tested. This chapter focuses on system models, training methods,

excitation signals, performance evaluation criteria and STARS implementation.

3.1 Aerodynamics System Model

The fundamental concepts of unsteady aerodynamics were used to develop a

generalized and sufficiently complex aerodynamic representation form. This required

determining the fundamental requirements needed for aerodynamic predictions of

aeroservoelastic systems.

3.1.1 CFD Solver

The objective of creating the aerodynamic system model is to replace the slow

CFD solver with a faster aerodynamics representation. The goal is to represent an

arbitrary CFD input-output relationship with an aerodynamic system model. In a general

sense, the source of the input and output is irrelevant, only the system representation

matters. CFD verification and validation remains with the CFD developer. The primary

function of the CFD solver for this thesis is to provide input-output vector pairs. The

CFD solution does not need to represent reality. However, because the CFD solver does

represent an actual physical system, the overwhelming problem of designing a general

34

system model is reduced. Regardless, the system model structure needs to be consistent

with the CFD solver. A review of the euler3d CFD solver’s solution structure is required.

Aerodynamic solutions are determined through time marching of a flow solver

with structural boundary conditions. Boundary conditions are specified through each

element’s displacement and velocity.

Structural response is determined through a discrete time representation of state

and input responses. In continuous time state-space form, the structural state matrices are

formed from the mass, damping and stiffness. When converted to a discrete time

representation, the state transition expression becomes the following:

[ ] [ ][ ] [ ] )()()(

)()()(tFDtxCtqtFBtxAtx

sss

ssss

+=+=&

It is important to note that the representation updates the state vector simultaneously.

3.1.2 Aerodynamic Specific Requirements

Aerodynamic systems have unique requirements when compared to other

systems. Together, these requirements force the reevaluation of common system

modeling techniques.

Aerodynamic systems have input and output dynamics. Fluids have memory of

both motion and flow geometries. This creates a situation where both motion and flow

must be adequately excited to capture the governing physics. Not exciting both

geometries will bias the ARMA model towards unrealistic coefficients for the motion and

force terms. These fluid geometries often have vastly different time-scales. Wake

convection off a body occurs at the free stream velocity, and wave propagation occurs at

sonic velocity. Worse still, wave propagation at transonic Mach numbers has vastly

35

different upstream and downstream clearing times. The aerodynamic modeling technique

must excite these dominant terms.

Aerodynamic systems have practical limitations on motion forms and magnitudes.

Exceeding these limitations places the input-output physics outside the linear regime. For

motion magnitudes, flow separation often effectively determines the linear limitations.

Additional limitations are placed on motion forms due to restrictions on starting and

ending input conditions. Any motion should begin from a steady state aerodynamic

solution to prevent generating unwanted impulsive transients. Steady state implies that

the entire flow field is constant. Steady forces are necessary but not sufficient for a steady

state condition.

Another requirement is that the aerodynamic solution and the boundary conditions

are consistent in a discrete time sense. This requirement becomes more critical with the

current displacement-based ARMA formulation. The training signal needs the same

motion characteristics that the coupled CFD solver uses.

Performance requirements for aeroelastic systems are significantly stricter than

the typical system identification requirements. The aeroelastic geometries under

consideration often operate over wide ranges of dynamic pressure. More seriously, the

prediction point of interest is often the system stability point. As Cowan [2003] found,

stability predictions are sensitive to solution technique.

3.1.3 Objective Function

This section describes unsteady aerodynamics as a generalized objective function.

It seeks to compare the unsteady forms and to select a viable representation, which will

be usable for later description-function selection.

36

Unsteady aerodynamics are complicated. Selecting an objective function based on

a linear representation is not trivial. Determining an objective function directly from a

fluid flow governing equation is just not possible.

The first step in selecting an objective function is to select the physics of the

unsteady flow that should be modeled. An initial selection criterion is to discard any

physics that are not included in the CFD solution. Since the CFD solver is an inviscid

Eulerian solver, this removes non-linear effects due to viscosity and fluid particle

tracking. Another selection criterion results from the use of structural modes for

aerodynamic boundary conditions. This allows only tracking the modal forces and

motions in lieu of tracking the entire flow solution. Now, the aerodynamic modeling

requirements are reduced to a MIMO system for predicting time marched modal forces.

Next, additional assumptions are placed on the system via the expected

magnitudes and forms of the inputs and outputs. The aerodynamic system may have a

static offset; this is easily removed. For most analytic solutions, including aerodynamics,

the residuals remaining, after assuming a strictly linear relationship between current

inputs and outputs, become sufficiently small for small enough inputs. However, this is

an unusually restrictive assumption. Further assumptions need to be based on unsteady

aerodynamic theory.

A review of unsteady aerodynamic theory with respect to input-output

relationships assists in determining a representative objective function. Reviewing the

Theodorsen solution to a harmonically moving airfoil shows that a frequency dependent

phase shift occurs for the non-circulatory portion of the lift. The Theodorsen solution also

indicates that motion displacement, velocity, and acceleration are included in

37

incompressible unsteady aerodynamics. From the wave equation for compressible flow,

additional unsteady effects will occur due to wave propagation resulting from changes in

boundary conditions and vorticity-boundary interactions. Complicating these effects is

that the waves themselves are affected by compressibility. For small amplitude motions

around a mean operating condition, the wave magnitudes are small. A final unsteady

effect occurs due to time lagged responses between two separated bodies. Time lags

occur at wave, potential and convective time scales. From system theory, pure time lags

result in distributed systems. That is, the continuous time transfer function is irrational

[Chen, 1999]. Further investigations yield even further complexities; however, it is fair to

say that the objective function needs to capture magnitude and phase changes with

frequency and also needs to capture pure time lags.

The function should model the dominant flow physics with a linear relationship in

the differential equation sense. Selecting a function based on an m by n order differential

equation provides an initial starting point. The following form is proposed:

)()( )()( delaysxxxxxcdelaysfffff nm ++++++=+++++ L&&&&&&L

This form captures the frequency dependent magnitudes and phases as well as any delay

terms. Additionally, the input dynamics and the output dynamics are observable within

the function.

3.1.4 Description Function Selection

The ARMA formulation is retained as the description function. The ARMA

formulation is already implemented in STARS.

The ARMA form allows for arbitrary high order descriptions through discrete

derivative approximations. The literature review indicated that the indicial forms were

38

not capable of properly modeling the boundary conditions. It was shown that the ARMA

formulation is capable of directly relating a continuous time ODE to a discrete form.

The ARMA form allows for both input and output dynamics. Correctly modeling

both the input and output dynamics is important as shown with the Theodorsen arbitrary

motion derivation. It appears that sensitivities in the current ARMA form are coming

primarily from excitation signal sensitivities.

In summary, the ARMA form allows for aerodynamic prediction. The ARMA

form is retained as the system model description function.

3.1.5 ARMA Realizations and Canonical Forms

There are an infinite number of realizations to describe an ARMA model. Each

form can be expressed as another through a specific transformation. The flow diagram for

the classical ARMA description is shown in Figure 3.1. The ai and bi terms are switched

when compared to the traditional ARMA naming convention. This form corresponds with

the formulation given in section 2.6.2. The past inputs and outputs are stored as separate

states.

One possible canonical ARMA form is shown in Figure 3.2. This formulation has

the past input/output information stored as one state, w(k). This form gives the advantage

of fewer states with the disadvantage of non-intuitive states. The reduction in the state

vector size is the minimum value of either na or nb.

39

Figure 3.1 Classical ARMA Form [Boziac, 1979]

Figure 3.2 Canonic ARMA Form [Boziac, 1979]

The state space representation for the canonical form is derived below. From

Figure 3.2, a logical state vector would be:

−−

=

)(

)2()1(

)(

nakw

kwkw

kxaM

A state update expression can be realized from Figure 3.2. However, it should be noticed

that Figure 3.2 swaps the ai and bi coefficients when compared to the traditional ARMA

convention used in this thesis. After substitution, the following state update and output

expressions are determined as the following:

)()()1( kxHkxGkx aaaa +=+

)()()(ˆ kxDkxCkf aaa +=

The canonical state matrices are:

40

−−−

=

010

00111

L

MMOM

L

L nana

a

aaa

G

=

0

01

MaH

[ ]nanaa abbabbabbC 0202101 −−−= L [ ]0bDa =

Comparison with the classical ARMA formulation given in section 2.6.2 shows that this

canonical formulation is a drop-in replacement.

To investigate the characteristics of the classical ARMA form and this canonical

ARMA form, system properties for a simple system are evaluated with the following

system:

∑∑==

−⋅=−⋅−4

0

4

1)()()(

ii

ii ikxbikyaky

The system has a model order of 4-5. There are 4 past force coefficients and 5 past

motion coefficients.

With the classical ARMA form, the aerodynamic system is the following:

−−−−

=

0100000000100000000100000000000000000100000000100000000143211111 bbbbaaaa

Ga

=

00010000b

H a

[ ]43214321 bbbbaaaaCa −−−−= [ ]0bDa =

From system theory [Chen, 1998], a system’s transfer function in terms of the state space

matrices is the following:

41

( ) DBAsICsG +−= −1)(

Thus, the investigated system has the following transfer function, which is identical to the

ARMA transfer function given in the previous section:

41

32

23

14

41

32

23

14

0)(azazazazbzbzbzbzbsG

++++++++

=

The canonical ARMA form is investigated next. The following state space

matrices are determined:

−−−−

=

010000100001

4321 aaaa

Ga

=

0001

aH

[ ]404303202101 abbabbabbabbCa −−−−= 0bDa =

Again, the transfer function exactly matches ARMA theory:

41

32

23

14

41

32

23

14

0)(azazazazbzbzbzbzbsG

++++++++

=

A slightly subtle result occurs when the eigenvalues of the classical aerodynamic

plant matrix are determined. The nb terms only contribute to the zero valued eigenvalues.

As a result, the classical and canonical forms contain the same eigenvalues.

The transfer functions for the classical, canonical and theoretical ARMA forms

are identical. For memory usage, the canonical form is preferred because the number of

system states is reduced to the maximum number of states of either the output AR terms

or the input MA terms. For eigenvalue analysis, the canonical form is preferred because

the plant matrix contains fewer terms. However, a major advantage that the classical form

offers is an intuitive state vector. The canonical state vector contains non-intuitive

42

transformed states. The output matrix Ca for the canonical form displays the decreasing

influence of past states in a more intuitive manner than the classical ARMA form.

3.1.6 ARMA Model Transfer Function

Describing the ARMA model as a discrete time filter provides another viewpoint

of the ARMA performance characteristics. From Boziac [1979], the z transfer function

for an ARMA model with model order na-nb is the following:

=

=

+= na

m

mm

nb

n

nn

za

zbzH

1

1

0

1)(

The ARMA model coefficients directly specify the transfer function coefficients.

3.1.7 ARMA Pathology

The ARMA system form is complex enough to allow strange, tricky and

inconsistent behavior for certain conditions. This section describes a few of these non-

intuitive situations, which might be useful.

3.1.7.1 Non-Consistent Training

An ARMA model is capable of describing a wide range of systems; however, the

model quality and coefficient form depend on the input training conditions. With the

current form, only input displacements are correlated with outputs. Training with a

velocity and no resulting displacement will associate the dynamics solely with the past

forces. The only terms that should be non-zero are the past force coefficients, nb terms.

This situation is more common than typically thought. If the motion update expression

43

does not update both displacement and velocity states simultaneously, an off-by-one-

timestep problem can excite this non-consistent training error.

There is a specific case where non-consistent training could be advantageous.

Combining inconsistent boundary conditions with an ARMA model that captures both

boundary conditions would allow training for each boundary condition for each mode.

This further decoupling of aerodynamics might yield better models; however, significant

challenges would result from combining all of the boundary condition models together

for each mode. Additionally, the ARMA form would require extra states to accommodate

the two boundary conditions. Combining the states together in a physically consistent

manner is not trivial. To examine this training method, an initial test was conducted by

adding a single “current” velocity state to the ARMA formulation. Velocity information

is determined from displacement data, so a possible half-timestep error is possible with

the displacement only ARMA form. However, the resulting models for the single velocity

addition were not significantly better.

3.1.7.2 Second Order Accurate Forces

Through a simple modification of the output matrix for the ARMA system, the

output forces can be described by a second order accurate description. The general

second order description for current force is:

)2()1()(2 21

23

2 −+−−= nfnfnff nd

Understanding that the ARMA model structure contains previous forces allows for the

introduction of the second order correction terms directly into the ARMA output matrix

structure. The general non-dimensional force is:

0)()(ˆ fkqDkxCf aaaa ++= .

44

Adding the 2nd order corrections terms gives the following force equation:

0221

123 )()()()(2ˆ fkqDkxSkxSkxCf aaaaaa +++−=

The S correction terms are:

[ ] [ ] [ ][ ][ ] [ ] [ ][ ]00

00

1

1

L

L

ISIS

==

Thus, the overall system output matrix is:

[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ][ ]1221141

243

22 −−−+−= nbnbnanaa BBBBAAIAIAC LL

Unfortunately, this derivation is not applicable for na < 2. This could be remedied by

adding two special “past force” states to the state vector.

3.2 Training Method

The training method concerns how a system identification routine generates a

system model from raw input-output data. For aerodynamic system identification, this

involves reducing aerodynamics force histories to usable aerodynamics system models.

The ultimate objective is to arrive at a useful system model with desirable characteristics.

This section will examine the data reduction, data flow and training methodologies used

to generate system models.

3.2.1 System Identification Data Flow

System identification data flow concerns how the raw aerodynamic training data

is transformed into an aerodynamic system model. Figure 3.3 plots the data flow from the

training signal xn.dat file to the system model file, .mdl.

45

Readxn.dat Detrend

SVD oneach mode

Packageinto .mdlFA

XN

FA

OFFSETS

A, B

Figure 3.3 System Identification Flow

The raw aerodynamic training data comes from an xn.dat file in the euler3d file format.

Raw aerodynamic forces are the generalized forces that are calculated inside the CFD

solver. Refer to section 1.3 for technical details of generalized forces. These raw

aerodynamic forces, FA, are de-trended based on the steady state values before an input

signal is applied. The XN and FA terms are combined into a matrix for each mode. SVD

determines the system coefficients for each mode. These system coefficients are then

packaged into the model file, .mdl.

3.2.2 SVD Data Flow

This subsection describes specifics of the data flow fed into the SVD routine. The

objective is to perform system identification on each mode and return the appropriate

ARMA model coefficients. The general data flow is shown in Figure 3.4.

Delayednb terms

Delayedna terms

CurrentForces

SVD Back SubstitutionSVD

FA

XNC

C

D

Repeated for each mode

Bi

Ai

Figure 3.4 SVD Data Flow

The XN data comes directly from the motion history. The FA data has been detrended as

discussed in the previous section. The delayed nb terms, which are calculated from the

46

motion terms, remain constant throughout the entire identification process. Data flow into

the SVD routine requires an ordered input and output relationship based on the ARMA

model form. For each mode, the delayed force terms and the delayed motion terms are

input into the SVD solver. Back substitution with the current forces yields the model’s

delay coefficients.

3.2.3 Time Scales and Aero-Structural Integration

Compatibility is needed between the discrete time model, the governing physics

and the flow solver. The flow solver is responsible for ensuring an accurate simulation of

physics; however, the flow solution quality is typically determined by the solution time

scale. A smaller solver timescale results in higher resolution at the expense of solution

time. A similar situation occurs with the discrete time model and the CFD solver.

Because the model is based on a discrete time representation, and the model only

contains a finite number of terms, the compatibility between the model and physics is

governed by the model’s input/output window. Figure 3.5 shows the relationships

between CFD, model and physical time scales.

CFD Model Physical

Requires a small time scale for solution convergence

Needs a large ‘window’to see relevant physics

Requires a small time scale to capture physics

Figure 3.5 CFD, Model and Physical Timescale Relationships

The figure shows the timescale relationships for aeroservoelastic analysis. The upper path

indicates that the basic CFD solution needs a small enough timestep to converge the

47

numerical solution. Failure to converge the solution in time will result in a non-physical

CFD solution. For example, if a dominant aerodynamic response occurs at 10 Hertz, then

sampling at 1 Hertz is assured to misrepresent the actual system. The lower path indicates

the system model timescale requirements. Between the CFD solver and the model, a

small enough time scale is needed for the model to capture the dominant physics. This is

the system model analog of the upper path. Between the model and physical systems, a

large enough window is needed to see the relevant physics. This means that the system

model must store and use sufficient information to accurately represent the physical

system. Together, these timescale criteria establish that converging the system model will

will require small timesteps and many states.

A further complication arises during the discrete time integration of aerodynamic

forces and the structure. Aerodynamic forces are applied with a zero order hold during

the structural response calculations. The assumption of a constant force becomes less

valid as the discrete timestep increases. For aeroelasticity, this assumption causes a

numerical error that manifests itself as structural damping. This result can be qualitatively

verified by noticing that the error in force is proportional to the change in displacement.

This is also equivalent to the correct force being applied at the incorrect time. A simple

solution to this problem is to use a smaller timestep.

For the discrete time representation, the eigenvalues are plotted in the z plane.

The root locus plot for a discrete time system hinges on the Nyquist frequency. The

Nyquist frequency is defined as half the sampling frequency. The relationship between

the damped natural frequency, undamped natural frequency, Nyquist frequency and

48

damping ratio are visually shown for the z plane root locus in Figure 3.6. The root locus

is symmetric about the imaginary axis.

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

Re(z)

Im(z

) 0.2

0.4

0.6

0.8

1 0

0.1

0.2

0.3

0.40.5

0.6

0.7

0.8

0.9

damping ratio ζ

0.0

Natural FrequencyRatio to Nyquist

Damped FrequencyRatio to Nyquist

Nyquist

Figure 3.6 Discrete Time Root Locus

Eigenvalues for aeroelastic predictions are concentrated below 20 percent of the

Nyquist frequency. Eigenvalues corresponding to the structural free vibration frequencies

are typically below 10 percent of the Nyquist frequency.

3.2.4 Training Data Redundancy

System identification requires associating the inputs and outputs. An important

consideration in developing a training method is determining the training signal duration.

The amount of information the system identification routine has available determines the

system model’s quality. From the ARMA model theory, the number of coefficients, N,

required for a model of order na-nb with nr modes is:

2),( nrnbnrnanbnaN ⋅+⋅=

The number of coefficients essentially scales with the number of modes squared.

Multiple mode and coupled systems require significantly more parameters because each

mode’s force depends on the inputs of all other modes.

49

Returning to the fundamentals of algebra, a unique solution for P unknowns

requires an equal number of equations. Therefore, training an ARMA model requires at

least N(na,nb) data points. It is important to understand that fewer data points will still

provide a model; however, the underdetermined model coefficients will be selected based

on the data reduction routine and not the training signal information. The resulting model

will predict the output based on the training signal, but it might not properly predict the

output for any other signal.

In practice, over-specification is required to maximize model performance. So,

the number of points needed for a model of order na-nb with an over specification factor

of F percent is:

( )100

),,( 2 FnrnbnrnaFnbnaN ⋅+⋅=

Over-specifying gives higher quality models at the expense of longer training signals.

3.2.5 Serial Training

Series training excites all modes sequentially during one simulation. This is the

traditional approach to system identification. Excitation of each mode is staggered so that

the system identification routine can distinguish between the individual input responses.

Series training offers the advantage of training all modes during one data capture.

Unfortunately, this data capture method allows for no errors. If the training fails during

the last mode, the entire signal is wasted.

3.2.6 Parallel Training

Parallel training excites each mode separately during separate simulations. This

has the advantage of requiring multiple short data captures. Thus, multiple simulations

50

can be performed at the same time. Testcase completion time is now proportional to the

number of modes and the number of computers.

Multiple small jobs are easier than one long job. Parallel training has a better fault

tolerance. Additionally, the parallel training creates better models because the modes are

separated. Because only one mode is being excited, all effects seen in the time history are

caused by only that mode. There is no long-term contamination or force dispersion as

seen in the serial training method. Parallel training allows for better modal specifications.

If a problem exists for one mode, only that mode must be redone. Additionally, training

properties can be tailored for each particular mode without significant computational

bookkeeping.

A logical extension of parallel training is to use multiple clustered computers.

Boeckman [2003] provides a parallel processing investigation. Aeroelastic prediction

solution time decreases with the number of modes and the number of computers for the

parallel training methodology.

3.2.7 Model Splicing

Using parallel training means that the training signals need to be combined. The

possible methodologies used for signal combination vary in complexity. However, the

fundamental objective remains consistent. The input signals should be joined in a manner

that reflects the problem physics and model structure. The following methodologies are

all possible; however, the later methods are more elegant and precise.

51

3.2.7.1 Simple Splice

A simple method of joining nr training signals together is to simply paste the

signals back to back to form an equivalent serial training signal. This requires the training

signal forces to start and finish at steady state values of force and motion. Otherwise, a

unit step is imposed, and the system identification will associate the resulting unit step in

forces with some non-physical excitation. One workaround is to perturb the de-trended

forces to steady state values of zero. This partially violates linearity, but for small

magnitudes, the method is appropriate. Figure 3.7 shows the system identification data

flow for this simple splice method.

Readxn.dat Detrend

SVD oneach mode

Packageinto .mdlFA

XN

FA

OFFSETS

A, B

Splice

Figure 3.7 Splice System Identification Flow

The SVD routine eventually breaks the training signal apart anyways, so this

method is not a strictly logical solution. Another disadvantage is that this method requires

the output to achieve steady state. Depending on the problem and the convection speed,

the solution might take longer to achieve steady state than to train the entire set of modes.

The simple splice generates useful models, but the method has significant disadvantages.

3.2.7.2 Model Superposition

Another possibility is to generate a system identification model for each mode and

then superimpose the models. This would allow decoupled system model identification

possibilities. For multi-mode system, training each mode as a SISO system and then

52

combining the results would determine the diagonal influence coefficients. Computing

the off diagonal terms will require correlating one mode’s output with another’s input.

System identification applies to this case and gives accurate SISO system models. The

problem occurs when the models are combined. Redundant system coefficients occur for

the na past force terms because the ARMA system identification associates forces with

both the motion and the past forces. In theory, the past force coefficients should be

identical. Adding the motion coefficients together should generate a useful model. In

practice, the past force coefficients are not identical. Small changes occur in the system

identification when trained as SISO system, so a methodology must be devised to select

past force coefficients. At this point, the superposition technique falls apart.

Of the many possible selection methodologies available, two appear promising.

The first and most logical solution is to use the na terms from only the diagonal influence

coefficients. As expected, this distorts the off-diagonal predictions. The second solution

is to average the nb terms for a particular mode. This distorts all of the predictions.

In practice, superposition of model coefficients does not effectively preserve the

system dynamics. Significant computational and selection energy is required to determine

less accurate models. Worse still, this system identification technique ignores the innate

MIMO abilities allowed by SVD in the interest of a simple decoupled identification

routine. This method does not produce viable system identification models.

3.2.7.3 True Splice

A more logical method of splicing multiple time histories is to use a true splice

using system identification flow information. This method is possible by taking

53

advantage of the discrete representation in the SVD data structure. The data structure is

reproduced below. Each row corresponds to a single timestep.

}

−−−−−−

ForceOutput Current ntsDisplaceme

11

Forces Previous

)()1()1()()()()1(M

44444444444 844444444444 76

MLM

LLL4444 84444 76

M

L tfnbtqnbtqtqtqnatftfnr nrnr

The logical jump required is to realize that the data structure appears to be a time

history, but it is in actuality only an unusually formatted matrix. The rules of linear

algebra apply. Individual columns cannot be eliminated or rearranged because they

represent the form of the system model. However, individual rows may be eliminated or

rearranged. Thus, the true splice solution to the parallel splice problem is to simply

catenate the training data together row by row in the SVD data structure. The only caveat

is to ensure that previous forces and previous displacement occurring before time zero are

set to zero. Only one offset force vector is assured because the signals are starting from

the same initial conditions. This method assures that there are no modal contamination or

long-term dispersion effects. Additionally, the routine has no extra computations inside

the SVD routine.

3.3 Excitation Signals

This section concerns input signal design. For the aerodynamic identification

routine to capture a useful model, the dominant aerodynamics must be excited. The

excitation is directly related to the input signal. This section is divided into two major

parts: input signal criteria and input signal investigation. The input signal portion is

developed in parallel and can be reviewed piecemeal.

54

3.3.1 Criteria

Evaluating the input signal performance requires criteria based on signal

properties and overall system performance. The criteria developed in this section will be

used to evaluate the input signals presented later in this section. The criteria are based on

physics, performance, and system model characteristics. The criteria are listed below:

• Excitation Consistent with the Aerodynamic Objective Function

• Excite Dominate Aerodynamics while still being in Linear Range

• Input-Output Dynamics must be Excited

• Starting and Ending Conditions for Input

• Excitation and Identification Based on Displacement

• Consistent Discrete Time Boundary Conditions

• Sufficient Spectral Power

The first criterion is that the input signal must excite the system in a way that is

consistent with the aerodynamic objective function. The signal must allow the static

offset forces to be determined. Additionally, the signal magnitudes must be large enough

to excite the dominant unsteady aerodynamics while still being kept in the “linear”

aerodynamics range. The objective function also shows that both input and output

dynamics must be excited. In comparison with the Theodorsen unsteady aerodynamic

theory, excitation must excite both the non-circulatory and circulatory lift in a way that

allows the system model to distinguish between the two.

A starting condition restriction is imparted on the excitation signal based on

physics. The input signal must start from rest and have a steady state flow solution to

ensure the unsteady solution accuracy. This condition significantly restricts the choice of

55

input signals presented in the literature. Specifically, this restricts input signals to motions

with either step changes, zero frequency starting conditions, or envelopes.

An additional criterion is placed on the excitation signal because the ARMA

representation form only uses modal displacements for input correlation. This means that

the excitation must train all input motion parameters based on displacement. Consistent

boundary conditions in the discrete time sense are extremely important. Since the goal of

the aerodynamic system model is to match the CFD solution, this boundary condition

criterion implies that the training signal should use the same motion update

characteristics as the coupled aero-structural system. This topic is discussed in more

detail in section 3.3.12.

Spectral power requirements set another input signal criterion. The input signal

must excite the aerodynamic system with sufficient power over a useful frequency range.

The power spectral density of the signal should be smooth across a specifiable frequency

band. At this point, a significant annoyance occurs. From system theory, integration

transforms a hypothetical “flat” PSD signal to a decreasing PSD with a slope of -20 dB

per decade. This causes difficulties. Specifying a “flat” PSD for displacement yields an

input signal with poor low frequency velocity information and even worse low frequency

acceleration information. It is necessary, but not sufficient, to base input signal PSD

criteria on only one motion specification. Based on physics, it appears that the “lower

derivative” terms become less important as frequency increases.

3.3.2 3211 Multistep

The 3211 multistep is the current training signal based on Cowan’s thesis work

[1998]. The signal consists of a unit step of lengths 3, 2, 1 and 1 unit implemented on

56

velocity. Figure 3.8 shows the displacement and velocity for the multistep. Displacement

is determined through numerical integration of the velocity. A dc offset occurs in the

displacement because the velocity signal is not symmetrical. The multistep’s nominal

length is 7 units long. The relationship between the unit length and a true time length is

specified by isize.

−2 0 2 4 6 8 10

−1

−0.5

0

0.5

1

Velocity

−2 0 2 4 6 8 10

−1

−0.5

0

0.5

1

Displacement

Figure 3.8 Multistep: Motion

A PSD for the multistep is given in Figure 3.9 for displacement and velocity.

Changing the isize unit length factor does not affect the PSD form, but does change the

overall power magnitudes. The displacement PSD displays the characteristic -20 dB per

decade decrease in power when compared to the velocity PSD.

57

0 10 20 30 40 50 60 70 80 90 10010

−4

10−2

100

102

104

Displacement

PS

D

0 10 20 30 40 50 60 70 80 90 10010

−2

100

102

104

Frequency (%Nyquist)

Velocity

PS

D

Figure 3.9 Multistep: PSD

The multistep has the following advantages. The 3211 multistep is easily

implemented due to the simple step changes in boundary conditions. The starting-from-

rest condition is automatically satisfied. Additionally, the maximum motion magnitudes

are describable in a closed form solution. The excitation frequency spectrum covers a

wide range with good low frequency content.

There are several disadvantages to the multistep signal. First, the multistep does

not excite the higher order terms in a useful manner consistent with the objective

function. This occurs because the multistep has step inputs in velocity. Acceleration

training data is limited to one timestep after each change in velocity. Any unsteady forces

due to acceleration terms will be captured in a non-physical manner. The second problem

with the multistep is that the PSD contains holes. This has the effect of not exciting those

particular frequencies and creating system models with poor transfer function properties.

58

A third disadvantage occurs because the velocity is strictly specified which causes

inconsistencies in boundary conditions due to non-simultaneous motion state vector

updates. A final disadvantage to the multistep is that the length sizing of the multistep

through the isize parameter is not intuitive. Isize is the critical parameter for correctly

exciting the dominant physics, but determining an optimal value for isize, given a

particular frequency range of interest, is not trivial. This analysis suggests that the

multistep should be rejected as an input signal for aerodynamic training.

3.3.3 Variable Amplitude Multistep

In an attempt to increase the 3211 multistep’s high frequency performance,

Cowan [1998] created the 753211 variable amplitude multistep (VAMS). Figure 3.10

shows the variable amplitude’s displacement and velocity.

−5 0 5 10 15 20 25−15

−10

−5

0

5

10

15Velocity

−5 0 5 10 15 20 25−1.5

−1

−0.5

0

0.5

1

1.5Displacement

Figure 3.10 Variable Amplitude Multistep: Motion

Because the variable amplitude multistep is a variant of the 3211 multistep, the

same disqualifying characteristics remain. Additionally, the VAMS introduces another

significant problem due to the velocity step amplitudes. The VAMS tends to cause

“hammering” of the aerodynamic system. The individual velocity steps have such large

amplitude that the training solution becomes non-linear or unstable. The VAMS is

59

particularly susceptible to this problem because the velocity ratio of the “7 unit” step to

the “1 unit” steps is 14 to 1. The VAMS is rejected as an aerodynamic training signal.

3.3.4 Chirp

The chirp input signal belongs to the analytic class of input signals. After determining

that the multistep signals have poor motion and frequency excitation properties, a new

class of signals based on frequency sweeps was developed. These signals allow for

smooth transitions in a specific frequency range. Sweep rate and length pairs will be

constrained by Nyquist frequency and data generation requirements. Frequency sweep

parameter selection is discussed further in Appendix C.

The chirp form is based on a linear frequency sweep. Expressions for

displacement, velocity and acceleration are given below. The chirp is analytic

everywhere. The chirp form is:

( )( )

( ) ( )22

2

2

tωttωtatωttv

tωtd

sin4cos2)(cos2)(

sin)(

22ωωω

−==

=

A time history plot of the chirp’s motion for displacement, velocity and acceleration is

given in Figure 3.11. Velocity is scaled by ω-1 and acceleration by ω-2. Velocity envelope

magnitudes increase linearly with time; acceleration magnitudes increase with the square

of time.

60

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−1

−0.5

0

0.5

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−2

−1

0

1

2

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−4

−2

0

2

4

Displacement

Velocity/ω

Acceleration/ω2

d(t)=sin(ω t2)

v(t)=2 ω t cos(ω t2)

a(t)=2 ω cos(ω t2) −4ω2 t2 sin(ω t2)

Figure 3.11 Chirp: Motion

A PSD for the chirp is given in Figure 3.12. The chirp sweep length determines

the maximum excitation frequency. This particular signal in Figure 3.12 excites up to

20% of the Nyquist frequency. Because the displacement has constant magnitude, the

PSD is relatively flat for displacement and is sloped at 20 dB per decade for velocity.

61

0 10 20 30 40 50 60 70 80 90 10010

−4

10−2

100

102

Displacement

PS

D

0 10 20 30 40 50 60 70 80 90 10010

−2

100

102

104

Frequency (%Nyquist)

Velocity

PS

D

Figure 3.12 Chirp: Power Spectral Density

The chirp has the following advantages. First, the chirp is an analytic function.

This means that the chirp is consistent with the objective function and is capable of

exciting all aerodynamic motion terms. The chirp starts from zero motion due to zero

frequency at time zero. The chirp’s displacement PSD is approximately flat up to the

excitation sweep length. A final and significant advantage is that the chirp has an

intuitive frequency excitation description. The excitation occurs up to the excitation

sweep length, which is easily determined and specified.

The disadvantages of the chirp are concentrated into two areas. First, the low

frequency performance of the chirp is poor for displacement and even worse for velocity.

The worst PSD for the chirp’s velocity occurs at low frequencies. However, this area is

important for aerodynamic predictions. From the Theodorsen function, the largest phase

difference for unsteady aerodynamics occurs at low reduced frequencies. The second

62

disadvantage results from the analytic nature of the chirp. Since the chirp is analytic, the

system model only contains analytic motion information. Non-analytic motions may not

be accurately represented. The analytic motion differs slightly from the actual system. In

actuality, this difference is significantly smaller than with the pure numerical technique

used for the multistep. This topic will be discussed further in section 3.3.12.

3.3.5 DC-Chirp

To alleviate the low frequency problems inherent in the chirp, a new dc offset

chirp was developed. The result is a non-symmetrical signal with a similar form but with

improved performance when compared to the chirp excitation signal.

The dc-chirp’s form is based on the original chirp. The signal is a linear frequency

sweep from zero frequency. Careful inspection shows that the dc-chirp is a peculiarly

integrated original chirp. Like the original chirp, the signal is analytic everywhere. The

functional form is:

( )( )

( ) ( )22

2

2

tωttωtatωttv

tωtd

cos4sin)(sin)(

cos)(

22

21

21

ωωω

−==

+−=

Figure 3.13 shows the motion history for displacement, velocity and acceleration. The

displacement envelope is held constant. This results in a linear envelope increase for

velocity and a squared increase for acceleration.

63

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−1

−0.5

0

0.5

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−2

−1

0

1

2

Displacement

Velocity/ω

Acceleration/ω2

d(t)= −2−1⋅ cos(ω t2)+2−1

v(t)= ω t sin(ω t2)

a(t)= ω sin(ω t2) + 2ω2 t2 cos(ω t2)

Figure 3.13 DC-Chirp: Motion

A PSD for the dc-chirp is given in Figure 3.14. Again, the frequency is swept to

20% of the Nyquist frequency. Careful inspection shows that the low frequency power

for the dc-chirp is increased over the original chirp’s power. The velocity PSD has the

same undesirable form as the original chirp at low frequencies.

64

0 10 20 30 40 50 60 70 80 90 10010

−10

10−5

100

105

Displacement

0 10 20 30 40 50 60 70 80 90 10010

−2

100

102

104

106

Frequency (%Nyquist)

Velocity

Figure 3.14 DC-Chirp: PSD

The dc-chirp has the same advantages as the original chirp plus additional low

frequency displacement power due to the dc offset. The dc-chirp allows for a visual

determination of rate and displacement effects because the displacement motion is offset

and the velocity is symmetrical. The different motion forms allow for improved

distinction between the effects of displacement and velocity. Figure 3.15 shows a classic

example of this motion distinction.

65

0 50 100 150 200 250 300 3500

0.2

0.4

0.6

0.8Displacement

0 50 100 150 200 250 300 350−0.05

0

0.05Velocity

0 50 100 150 200 250 300 350−40

−20

0

20Forces

Zero

Figure 3.15 DC-Chirp: Motion Determination Example

The dc-chirp has many of the same disadvantages as the original chirp. Adding

the dc offset removed the PSD problem at low frequencies for the displacement;

however, little power is being applied to the velocity at low frequencies. A subtle second

disadvantage concerns the peak factor of the dc-chirp compared to the original chirp.

Because the new dc-chirp is non-symmetrical, the absolute forced magnitude of

displacement is only half that of the original chirp with a maximum displacement of

unity. Doubling the dc-chirp’s magnitude to achieve a similar power level pushes the dc-

chirp closer to the non-linear aerodynamics regime.

3.3.6 Fresnel Chirp

The Fresnel chirp is a linear frequency sweep with a displacement as the true

integral of the original chirp. The Fresnel chirp is expected to increase the low frequency

velocity power. The result is an excitation signal with no closed form expression but with

potential for better training performance.

66

The Fresnel chirp is an integrated form of the original chirp. The Fresnel chirp

still contains a linear frequency sweep and is analytic everywhere. As a practical

implementation issue, there are two forms of the Fresnel chirp. The two forms result from

the integration of either the sine or cosine function. The signal needs to start from rest, so

the Fresnel sine function, S(t), is chosen. The functional form is:

( )( )

( )2

2

2

tωttatωtv

tωtStd

cos2)(sin)(

sin)()(

ω==

== ∫

A motion history for the Fresnel chirp is given in Figure 3.16. An envelope

magnitude occurs for the Fresnel chirp’s velocity. As expected by integration theory, the

displacement tapers down in magnitude as frequency increases. A PSD for the Fresnel

chirp is given in Figure 3.17.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−1

−0.5

0

0.5

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.2

0.4

0.6

0.8

Displacement

Velocity

d(t)=S(t)= ∫ sin(ω t2)

v(t)=sin(ω t2)

Figure 3.16 Fresnel Chirp: Motion

67

0 10 20 30 40 50 60 70 80 90 10010

−4

10−2

100

102

104

Displacement

PS

D

0 10 20 30 40 50 60 70 80 90 10010

−4

10−2

100

102

Frequency (%Nyquist)

Velocity

PS

D

Figure 3.17 Fresnel Chirp: PSD

The Fresnel chirp has significant advantages over those of the previous chirps.

The largest improvement consists of the “flat” velocity PSD. Velocity power is still weak

for low frequencies, but an integrated dc-chirp cannot be implemented because

integrating an offset signal yields a diverging signal. Integrating a non-symmetrical

velocity signal will cause displacement peak factor problems. A minor disadvantage of

the Fresnel chirp is that the resulting function has no simple, closed form solution and

requires numerical integration for practical implementation [Zwillinger, 1996]. The noisy

displacement PSD appears to be a result of the integration scheme introducing noise into

the displacement signal. Noisy integration may become a sensitivity issue later.

68

3.3.7 Schroeder Sweep

The Schroeder excitation signal is based on the multisine class of signals. This

class of signals allows for the specific excitation of a specified bandwidth through the

summation of discrete frequencies. A phase-shift presented by Schroeder [1977] allows

for minimizing the signal’s peak factor.

The Schroeder form is based on a sum of cosine terms with a specified phasing.

Expressions for displacement and velocity are given below. The form is analytic in time

but not smooth in frequency.

=

=

−−=

−=

N

k

N

k

Nk

Ttk

NTktv

Nk

Ttk

Ntd

1

2

1

2

2sin212)(

2cos21)(

πππ

ππ

Figure 3.18 plots the displacement and velocity for a 50-term Schroeder sweep.

The signal is harmonic and can continue indefinitely. The signal resembles a frequency

sweep during a significant portion of the time history. The Schroeder sweep appears to be

the discrete time counterpart to the continuous-time frequency-swept chirps. The

displacement is bounded by a unit magnitude envelope.

69

0 5 10 15−1

−0.5

0

0.5

1Displacement

0 5 10 15−30

−20

−10

0

10

20

30Velocity

Figure 3.18 Schroeder Sweep: Motion

A PSD for the Schroeder sweep is given in Figure 3.19. The number of terms in

the summation determines the sweep bandwidth. The PSD is flat along the excitation

bandwidth. This particular signal excites up to 20% of the Nyquist frequency. Careful

inspection shows that a pure dc component exists. The sweep is implemented on

displacement, so the velocity PSD exhibits the expected sloped response. However, the

Schroeder sweep could be implemented on velocity with only minor changes.

70

0 10 20 30 40 50 60 70 80 90 10010

−6

10−4

10−2

100

102

Displacement

PS

D

0 10 20 30 40 50 60 70 80 90 10010

−2

100

102

104

Frequency (%Nyquist)

Velocity

PS

D

Figure 3.19 Schroeder Sweep: PSD

The Schroeder sweep has the following advantages. First, the PSD is specified in

an intuitive manner and is flat. The Schroeder sweep at zero frequency has desirable

power levels for the excitation motion. Second, the Schroeder sweep has an optimal peak

factor when compared with both discrete and continuous “sweep” signals. For example, a

comparison between the Schroeder sweep and the corresponding sinc function is shown

in Figure 3.20. Both signals have the same PSD characteristics. The Schroeder sweep

shown at the bottom of the figure has a peak factor 4 times smaller than the constant

phase multisine at the top of the figure. The constant phase multisine is also equivalent to

the sinc function.

71

-1

0

1

0

1

2

3

4

N

kk

2

1

πθθ −=

∑=

+=

N

k

kk

T

tkptr

1

2cos

2)( θ

π

2

πθ =k

Phasing:

Basis Function with N=30

Npk /1= (equal power at all frequencies)

Figure 3.20 Schroeder Sweep: Peak Factor Comparison

The disadvantages are concentrated in three areas. First, the Schroeder sweep

does not intrinsically contain a from-rest starting condition. Any attempt at using the

Schroeder sweep requires a modification to establish both the steady state system

behavior and the from-rest starting response. Bounding the sweep with an envelope will

cause undesirable and non-intuitive distortions of the signal’s characteristics. Simon

[2000] shows the undesirable results of harmonic signal truncation. Testing with the

Schroeder sweep verified this sensitivity. Second, the Schroeder sweep requires a

numerical sum of terms at each timestep. This summation is expensive for large

excitation bandwidths. The third disadvantage is that the Schroeder sweep, being a

harmonic signal, appears to be excessively sensitive to excitation length. Testing shows

that small errors in signal specification can cause regular holes in the Schroeder’s PSD.

These types of sensitivities are not desired for robust training.

72

3.3.8 Noise Training Signal

A final class of signals is based on stochastic system theory. White noise is

defined as a signal with equal power at all frequencies. A white noise input signal would

allow for equal excitation up to the Nyquist frequency. A white noise signal with 500

data points is given in Figure 3.21.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−1

−0.5

0

0.5

1

Figure 3.21 Noise: Displacement

A PSD for white noise is given in Figure 3.22. This plot uses 500 samples. A

second PSD for white noise with 5000 samples is given in Figure 3.23. Excitation length

strongly determines the PSD distribution characteristics. A constant PSD over the entire

frequency range requires an infinite excitation length.

0 10 20 30 40 50 60 70 80 90 10010

−3

10−2

10−1

100

101

PS

D

Figure 3.22 Noise: PSD 500 Samples

73

10−6

10−4

10−2

100

102

PS

D

Figure 3.23 Noise: PSD 5000 Samples

The white noise input signal has advantages. First, the PSD is defined to be “flat”.

The signal can start from rest and has no restrictions on excitation length. Unlike the

analytic signals, the white noise input will excite the non-harmonic components in the

aerodynamic system.

There are numerous disadvantages to the white noise signal. First, the “flat”

power level is only achievable as the excitation length becomes infinite. Therefore, the

excitation length is non-intuitive. The implementation question becomes: How long is

long enough? A related disadvantage is that the training signal is not deterministic. The

signal is unknown random noise. Traditional statistical data allows for comparisons

between the noise properties but does not allow for input-output comparisons. The noise

input signal is not guaranteed to excite the same system dynamics between two otherwise

identical simulations. Comparison of non-deterministic training signals is likely to

encounter difficulties. Second, the noise input signal creates problems with boundary

conditions in two ways. The first is that the input excitation is not consistent with the

discrete-time solution method. Equal power at the frequencies near Nyquist implies that

“sharp” changes occur in the input signal. These changes will overdrive the flow solution

solver. The motion time history in Figure 3.21 shows this characteristic of the signal. An

example of this problem would be an excitation that moves an airfoil so quickly that the

74

flow solution becomes dominated by shock acoustics, which violate the linearity

assumption. The second boundary condition problem is that simultaneous updating of

motion states is not possible. Because the signal is not deterministic, generating

displacement from velocity, or vice versa, requires numerical integration. Numerical

integration introduces time lags and is not consistent with the ARMA motion

specification. These disadvantages require that the noise signal be rejected as an input

signal.

3.3.9 Envelopes

A complication arises with some signals because their harmonic form does not

allow initial starting conditions without impulsive transients. A general method of either

beginning or ending the excitation when the excitation does not begin at zero is needed.

One solution is to develop an envelope term that starts or stops the input signal from a

zero motion condition.

Combining two exponential functions allows for a generic envelope function. An

envelope function based on this concept is given below. The power constant of 11 is

empirically determined based on function roll-off requirements. Higher power constants

yield sharper roll-offs. The envelope form is:

110

110 )exp()exp()( tttttenv −⋅−−=

While the above form appears only useful for ending conditions, the following

transformation allows for specifying a zero motion starting condition:

)(1)( tenvtenvstart −=

The above forms are poorly suited for numerical evaluation. A better numerical

expression is found by expanding into the following form:

75

( )100

830

650

470

290

110 223309246601102exp)( ttttttttttttenv ⋅−⋅−⋅−⋅−⋅−⋅−=

The effective time of roll-off is determined through the t0 term. Practically, this term

must be determined empirically; a general solution is not intuitive. However, a more

powerful technique is to non-dimensionalize the envelope time terms and then re-

dimensionalize the envelope for the particular timescale. Setting t0 equal to 1.0-10

specifies an envelope with a magnitude of approximately 0.1% of the initial value at a

non-dimensional time of unity. Figure 3.24 shows the time form of the envelope.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.2

0.4

0.6

0.8

1

Non−Dimensional Time

Envelope Magnitude

Figure 3.24 Envelope Form

3.3.10 Superposition of Multiple Signals

The objective is to supplement the regular training signal by adding extra signals

through superposition. Adding signals together allows for the possibility of exciting

multiple frequency ranges without using a single long sweep. However, this type of

superposition assures phase and frequency interference because the training signals have

restrictions on the starting conditions. As a result, the power spectral density has sharp

holes at the interference frequencies. Figure 3.25 shows the PSD of a superposition of a

chirp signal, a low frequency chirp signal and the resulting summed signal. The peak

factor is held constant.

76

102

10 1

100

10 1

100

101

102

Chirp Signal

Low FrequencyChirp

Sum

PSD

Frequency Figure 3.25 Superposition of Multiple Signals: PSD Holes

Sharper holes are created in the summed signal. This example shows the importance of

both magnitude and phase in signal superposition. These PSD characteristics of summed

signals show the difficulties of signal design with respect to peak factors. Choosing a

globally optimal signal appears to be better than choosing and combining several locally

optimal signals.

3.3.11 Purposefully Added Noise

Contrary to intuition, adding noise to an existing signal is often desired for system

identification. This occurs because many signals, like the chirp, only excite a specific and

limited frequency spectrum due to simulation time and aerodynamic convergence

concerns. This causes problems when the system identification correlates small

perturbations in the input signal with numerical artifacts in the output signal. The

resulting relationship between input and output becomes dominated by numerical

artifacts rather than the underlying physical system. Viewing the problem from a transfer

function approach yields the same conclusion. A ratio of output to input is not specified

77

by the absolute magnitudes of input and output. This situation occurs regardless of the

identification method used; it is the result of the training method. Specifically, the

numerical convergence of a frequency power level must be consistent between the input

and output. For strictly defined functions, the input noise is on the order of the storage

numerical accuracy. This problem typically occurs at high frequencies where the training

signals have little or no power. The problem surfaces as spurious and non-physical

eigenvalues in the complex z plane.

One solution is to add artificial noise to cover the high frequency spectrum. White

noise offers a wide frequency excitation when given sufficient excitation length. The

noise only needs to increase the input signal power levels; a corresponding increase in the

output is not necessary. To visually display this concept, a noise comparison experiment

was developed. The experiment used a subsonic two-dimensional airfoil with a single

pitch mode to generate the output signal. The input signal to the CFD solver was the

clean signal shown the lower plot of Figure 3.26. A noisy input signal was generated by

superimposing the clean signal with low amplitude white noise. The upper plot of Figure

3.26 shows this noisy signal. The same clean output signal was used for both the noisy

and clean systems. Thus, both systems use the same output with different inputs. The

noise was added to the input signal after the CFD solver but before the system

identification. This is only a conceptual experiment; an actual identification would not

use this ad-hoc noise augmentation method. The objective is to show the effect of signal

spectral content without introducing CFD related uncertainties and complications.

78

0 0.5 1 1.5 2 2.5 3 3.5

0

5

10

Noisy

1

0 0.5 1 1.5 2 2.5 3 3.5

0

5

10

1

Clean

Figure 3.26 Noise Experiment: Noisy and Clean Input Signal

A transfer function comparison is made from the raw input and output signals. To

achieve a fair comparison of the clean and noisy systems without generating intermediate

system models, the transfer functions were generated as the ratio of power-spectral-

densities of the output over the input. Figure 3.27 plots the frequency dependent transfer

function for the clean signal. Figure 3.28 plots the transfer function for the noisy signal.

For the clean signal, the transfer function form appears smooth; however, the

form does not agree with either intuition or physics. Regardless of what the system

transfer function indicates, forcing the system at the Nyquist frequency should not result

in a factor of 1020 more output than the steady state response. Output power converges to

a PSD level of approximately 10-13 at high frequencies due to the CFD solution

limitations. Yet, the input signal continues to decrease in power level as frequency

increases because the input signal is strictly defined as a mathematical function. Forcing

more noise in the input signal will assist finding a better system transfer function.

For the noisy signal, the transfer function does not appear smooth because the

white noise excitation is limited in length. However, the ratios of input and output powers

appear more consistent. The transfer function captures the dominant low frequency

structure of the aerodynamics without an overwhelming response at high frequencies.

79

0 10 20 30 40 50 60 70 80 90 10010

30

10 25

10 20

10 15

10 10

10 5

100

105

1010

1015

Transfer Function

Output

Input

Frequency (% Nyquist)

Power Spectral Density

Figure 3.27 Noise Experiment: Transfer Function with a Clean Input Signal

0 10 20 30 40 50 60 70 80 90 10010

14

10 12

10 10

10 8

10 6

10 4

10 2

100

Power Spectral Density

Transfer Function

Output

Input

Frequency (% Nyquist)

Figure 3.28 Noise Experiment: Transfer Function with a Noisy Input Signal

A second comparison is made with the coupled aerostructural system’s

eigenvalues. For this comparison, the low frequency eigenvalues inside the input training

signal’s main excitation sweep are ignored. The eigenvalues under consideration are

those higher than approximately 10 percent of the Nyquist frequency. Figure 3.29 plots

the eigenvalues in the z plane for the clean input signal. Undesirable eigenvalues at high

frequencies emanate from the origin as dynamic pressure is increased. These correspond

to the large magnitudes of the transfer function at high frequencies. Eigenvalues for the

noisy input signal are plotted in Figure 3.30. In contrast to the clean signal, the high

frequency eigenvalues for the noisy signal remain near the origin.

80

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 3.29 Noise Experiment: Eigenvalues with Clean Input Signal

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 3.30 Noise Experiment: Eigenvalues with Noisy Input Signal

Conceptually, the addition of high frequency power to the input training signal

improved the system identification model’s accuracy and performance. For high fidelity

simulations such as controls, the clean signal would cause significant simulation

problems, while the noisy signal would allow simulations that are more precise,

especially at high frequencies. It appears that storing and changing the input and output

vectors should be numerically consistent.

3.3.12 Motion Specification

This section investigates input signal motion specification. Two choices exist for

specifying the input training motion: strict motion specification and state-space motion

specification.

It is important to understand why motion specification is important. The ARMA

model associates forces with past and previous displacements and past forces. Only

through consistent boundary conditions is the ARMA form able to “see” velocities and

higher derivatives. The result is that the system identification routine requires a training

81

method consistent with the boundary conditions and the motion state update

methodology.

3.3.12.1 Strict Motion Specification

Strict motion specification is the current methodology. For strict motion

specification, the motion state vector at each timestep is determined from either an

analytical or a numerical expression without reference to the previous motion state

variable. For analytical input signals, this method results in “perfect” signals with

perfectly correct integrals and derivatives. For example, the dc-chirp and its derivative

could be used to directly determine the boundary conditions at any timestep.

The advantage of strict motion specification is that the implementation is simple.

The method only requires directly setting the boundary conditions inside the CFD solver.

The CFD flow solver does not care if the boundary conditions make no sense; the flow

solver will still output a solution.

The disadvantage of strict motion specification is important. Strictly specifying

the boundary conditions requires consistence in the discrete time sense. The method

requires either a continuous time derivation or a discrete time numerical operation.

Neither is correct.

3.3.12.2 State Space Motion Specification

An alternative method for motion specification is through a state space update

methodology. The boundary conditions are updated through a state space form with the

excitation signal not directly specifying the state vector. This is a methodology identical

with structural excitation with an external force function. Implementing this methodology

82

involves the derivation given in section 2.6.3. This methodology resembles the excitation

of a mechanical system without stiffness or damping.

The advantages of this methodology are significant. The boundary conditions are

updated simultaneously in a true discrete time sense. The excitation signal and response

is consistent with the coupled aero-structural response. Additionally, non-analytic input

signals are no longer a problem because the input is filtered through the “structure”. A

convenient advantage of this methodology is that the state space update routines already

exist in the CFD solver; there is no code development required.

The disadvantages of this motion specification methodology are slight. The

methodology requires selecting a force versus “structural mass” ratio that excites the

system within the linear limitations. Maximum magnitudes are not directly specifiable as

in the strict motion methodology. The biggest disadvantage is the stopping condition.

Because the excitation signal is input on acceleration, ending the excitation does not

result in the motion returning to zero. Instead, the mode will continue to move with

constant velocity after the excitation is removed. This can result in motions that exceed

linearity limitations. This requires an ending condition specification that requires the

velocity and displacement to remain within certain bounds. Practically, this can be

handled with a crude force feedback loop on displacement.

3.4 Model Performance Evaluation Criteria

This section seeks effective methods of comparing and selecting ARMA models

based on model performance. The fundamental difficulty is that all reasonable models fit

the training signal, so comparison and selection by visual inspection of time histories is

doomed to failure. However, any model not fitting the training signal is automatically

83

rejected. The preferred selection technique is to use a numerical measurement. Also, any

information on actual system model order is desired.

3.4.1 Chi Squared

Chi-squared error, 2χ , gives an indication of how well the SVD model fits the

training data. This is an indication of solution quality, not model performance quality.

The characteristics of the chi-squared error resemble the model fit least square error. Chi-

squared error is calculated with the following expression:

nstpbxA 2

2 −=

r

χ

The overall fit-error is divided by the number of time steps in the training signal, which

ensures a fair comparison between variable length signals. Otherwise, 2χ always

increases with an increase in data points for the same model.

Chi-squared error converges to zero as the number of model coefficients

approaches infinity. This is expected because more coefficients allow a closer fit

regardless of the resulting model quality. A typical plot of 2χ error and model order is

shown in Figure 3.31.

84

0 20 40 60 80 100 120 140 160 180 20010

−6

10−4

10−2

100

102

Number of Model Coefficients

χ2

Figure 3.31 Chi Square: Typical Two Mode Plot

The advantage of using chi-squared error as a selection criterion is that chi-

squared error directly relates to the solution quality. Large chi-squared errors indicate

poor model performance, which is caused by the model not accurately fitting the training

data. Global chi-squared magnitudes indicate the amount of noise contained in the

training signal. Another advantage is that determining chi-squared error is

computationally cheap; only one matrix multiplication is required.

The primary disadvantage of chi-squared error is that it is not useful for model

performance characterization. Since the chi-squared error always decreases for increasing

model order, it does not allow for model order estimation. Low chi-squared error is a

necessary, but not sufficient, condition for good quality aerodynamic models.

3.4.2 Force Prediction Root Mean Square

The force prediction RMS gives a measure of forced response prediction with a

known signal. Typically, the force RMS is applied to the training signal; however, any

signal is applicable. The force RMS calculation is made with the following expression

where )(ky refers to the predicted force vector and )(ˆ ky refers to the actual force vector:

85

( )∑=

−=N

kkyky

NRMS

0

2)(ˆ)(1

The force RMS provides significantly better model performance information than chi-

squared error because the force RMS system model contains aerodynamics indexed to the

starting conditions which are not reset to the known signal at each timestep. Unlike chi-

squared error, force RMS evaluates the true system performance for a connected time

history.

An example force RMS plot versus model coefficients for a two mode testcase is

given in Figure 3.32. For this testcase, the force RMS converges to near a minimum value

and then varies around a mean point. With further increases in model order, the RMS

slowly increases.

0 20 40 60 80 100 12010

−3

10−2

10−1

Number of Model Coefficients

Force RMS

Figure 3.32 Force RMS: Two Mode Testcase

While the force RMS does not appear to be an effective model order determination

criterion, it does give an indication of model quality.

3.4.3 Partial Autocorrelation

The partial autocorrelation correlates the response dependencies at each time with

all other times. Conceptually, PACF information could determine a limiting model order

86

that captures the dominant dynamics. The model order could be sized based on where the

PACF becomes effectively zero. This could potentially eliminate the current brute force

model selection methods. However, Brockwell [1991] points out a significant limitation

of the PACF:

In contrast with the partial autocorrelation function of an AR(p) process,

that of an MA(q) process does not vanish for large lags. It is however

bounded in absolute value by a geometrically decreasing function.

This indicates that the PACF would not provide accurate model order data for ARMA

models. For comparison, Figure 3.33 plots the PACF for an actual subsonic aerodynamic

system generated with CFD and a chirp training signal.

Figure 3.33 PACF: Aerodynamic System with Chirp Training Signal

The response curve of Figure 3.33 appears to agree with Brockwell’s prediction. No clear

minimum-model-order lag time stands out.

3.4.4 Coupled Aero-Structural Properties

A powerful model performance criterion is based on the coupled aero-structural

properties exhibited by the model. Using the model for coupled aero-structural

predictions exercises the model’s prediction characteristics in a realistic and intuitive

87

manner. This criterion allows for both individual and multiple model performance

evaluations.

This criterion is based on eigenvalue predictions. The eigenvalues are calculated

from a coupled aero-structural state transition matrix. Calculating eigenvalues

corresponding to the aerodynamic state-space model will only show the internal system

response and not the input-output response. This would be a major limitation for model

performance evaluation. The simple solution is to include structural dynamics. Using the

coupled aero-structural derivations given in section 2.6.3, a discrete time plant matrix is

determined. The expression is reproduced below:

+ ∞∞

asa

assass

GCHCHqCDHqG

Dynamic pressure, q∞, is an independent variable. Plotting the eigenvalues in the discrete-

time z plane over a range of dynamic pressures yields both stability boundary and model

sensitivity information. A typical root locus plot for a density sweep from 0 to 20 units of

dynamic pressure is given in Figure 3.34.

Figure 3.34 Eigenvalue Explanations

88

Because the plant matrix contains both the structural and aerodynamic systems,

eigenvalues corresponding to both systems appear. Structural motion eigenvalues begin

at the free vibration frequency for zero dynamic pressure. Eigenvalues corresponding to

the ARMA model are seen as the set of curves emanating from the z plane’s origin. The

model’s sensitivity to dynamic pressure and suitability to controls applications can be

determined from the eigenvalue locations and motion.

Evaluating multiple models allows for model order sensitivity studies. A typical

model order sensitivity plot is given in Figure 3.35.

0 10 20 30 40 50 60100

110

120

130

140

150

160

0

1

2

nb Order

Dyn

amic

Pre

ssur

e na Order

Figure 3.35 Model Order Sensitivity Explanation

For this plot, dynamic pressure for neutral stability is plotted for a range of na and

nb model orders. It is desired that the dynamic pressure for neutral stability converge

with both na and nb terms. Figure 3.35 is an example of poor convergence. Figure 3.36

shows excellent convergence.

89

0 5 10 15 20 25 300.4

0.45

0.5

0.55

0.6

0.65

0.7

0.75

0.8

1 23456789101112131415161718192

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 3.36 Model Order Sensitivity Convergence

A second type of eigenvalue sensitivity study is to plot the coupled aero-structural

eigenvalues for a range of dynamic pressures over a range of model orders. This study

indicates the degree of model prediction consistency. Typically, the eigenvalues of

concern for this type of study are those corresponding to the structural motion. Figure

3.37 shows a typical model sensitivity study with the coupled eigenvalues in the z plane.

The figure is zoomed to show only the structural eigenvalues. This particular testcase has

six modes.

Figure 3.37 Eigenvalue Sensitivity Study Explaination

90

For this example, the eigenvalues converged as the model order was increased

from 2-8 to 6-16. An order of 10-20 yields a slightly different eigenvalue structure. A

model order of 6-16 seems to be the best. This type of sensitivity study provides a

powerful method to evaluate the resulting aerodynamic system models.

The advantages of using coupled aero-structural performance evaluation criteria

are significant. First, coupling the two systems allows for evaluating a more complex

combined system. This combined system reveals problems more readily and tests the

system model in a realistic environment. This method allows for system property

evaluations that would be utterly invisible when comparing the responses in the time

domain.

There are disadvantages to using a coupled system performance criteria. The

method requires significant computational effort. Because the aerodynamic model order

changes, individual eigenvalue sweep comparisons for the aerodynamic terms are not

possible in the z plane.

3.5 Preliminary Testcases

These examples consist of simple testcases with known solutions. The objective is

to investigate system identification performance on simple testcases. This should assist

will interpretating and evaluating more complicated actual testcases.

3.5.1 Zero Order Force Function

This testcase consists of a forcing function that only depends on the current

motion. This results in a zero order differential equation: xqf &= . The resulting force is

coupled with a single degree of freedom structural system. The representation of the

91

overall structural system is fxkxcxm =++ &&& . From the simple force model, xqf &= , a

simple chirp training signal was used to generate an input motion versus output force

relationship.

Figure 3.38 Zero Order Force Training Signal and Output

From this training signal, three ARMA models were generated. Because the

current ARMA model structure stores only the displacements, the resulting ARMA

model should return the coefficients for a first order derivative. The model should contain

coefficients for a first derivative of equally spaced points with the derivative taken at the

endpoint. The following table shows the resulting non-dimensional ARMA model

coefficients and the exact coefficients from a Taylor series expansion. The ARMA

coefficients are non-dimensionalized by )!1(1 −ph , where p is the number of points and

h is the point spacing.

92

2 Pt 3 Pt 4 Pt

Step Taylor Series ARMA

Taylor Series ARMA

Taylor Series ARMA

n 1 1.007 3 2.9990 11 9.078 n-1 -1 -1.018 -4 -3.9981 -18 -12.235 n-2 1 0.9989 9 3.239 n-3 2 -0.082

χ2 0.39·10-2 0.29·10-5 0.28·10-5

Table 3.1 1st Derivative Comparison for Taylor series and ARMA coefficients

The coefficients from theory and ARMA agree for the 2 and 3 point derivative

expressions. The 4 point expression does not match other than qualitatively. However,

the 4 point derivative offers no better fit than the 3 point. This suggests that the model has

converged and extra data points are not likely to improve the data fit. The output fit

parameters χ2, RMS and autocorrelation (ACF) are plotted in Figure 3.39 for nb model

orders from 2 to 6. RMS and chi-squared errors level off at three terms. It appears that the

first bump of ACF corresponds to leveled RMS and chi-squared error.

Figure 3.39 Output Fit Parameters

Next, the entire system’s stability is investigated. The governing differential

equation is xqxkxcxm &&&& =++ . The system is expected to become unstable at cq >

93

when damping becomes zero. In state space form with the structural damping in terms of

the forcing function, the overall structural system becomes the following:

−−−=

xx

mqc

mk

xx

&&&

& 10

The system was discretized with the parameters given in Table 3.2:

Mass m 1.0Stiffness k 4.0Damping c 0.1Time Step ∆T 0.1

Table 3.2 0th Order Forcing Function Structural Parameters

The discrete time state space model becomes:

=

++

)()(

)1()1(

kxkx

Akxkx

d &&

The state transition matrix, Ad, is the following expression, which is in terms of structural

and aerodynamic parameters:

( )( )( )( )

−−⋅⋅−−

−⋅+⋅⋅=

xCqcSzzSzSqcSxCz

Ayyy

yyyd 2

2

The coefficients used above are factored using the following coefficients:

( )( )

)sin()cos(

)(exp

24

121

21

22

ySyC

qxtqcz

txyqcqckx

y

y

=

=⋅∆⋅−−=

∆⋅=−+−=

At zero dynamic pressure, q=0, the state transition matrix is the following:

=97.0395.0099.098.0

dA

94

Plotting the eigenvalues of this system over a range of dynamic pressures of 0 to 20

yields Figure 3.40 and the zoomed eigenvalue crossing Figure 3.41. The eigenvalues

cross the unit circle at a dynamic pressure of 0.1 as expected. At a dynamic pressure of

4.1, the eigenvalues meet at the real axis and the structure becomes statically divergent.

Figure 3.40 Zero Order Forces: Eigenvalues

Figure 3.41 Zero Order Forces: Zoomed Eigenvalue Crossing

Now, the coupled aero-structural eigenvalues are determined using system identification.

Figure 3.42 plots the eigenvalues for model orders from 0-2 to 0-7. The structural

eigenvalues qualitatively match the exact form shown above. All models determine a

stability boundary at a dynamic pressure of 0.1.

95

0-2 Model

0-3 Model

0-4 Model

0-5 Model

0-6 Model

0-7 Model

Figure 3.42 Zero Order Forces: System Identification Eigenvalues.

The extra aerodynamic states appear as eigenvalues moving away from the origin. The

frequency spacing of these extra terms are constant, which reveals the equally spaced

discrete-time form of the ARMA model.

3.5.2 Second Order Force Function

A more complicated example is created using the same structural system and a

more complicated forcing function. This testcase consists of a forcing function that

depends on a second order ODE. The differential equation for force is:

xxfff &&&& 1.082 +=++

f is an output and x is an input. Converting the system to a state space form with the

states [ ]Tfff &r

= yields the following:

[ ] [ ]xBfAxx

ff

ff rr

&&&&

&+=

+

−−

=

1.01

0028

10

96

For the continuous case, determining an analytical stability boundary is possible

by coupling the structural and forcing systems and evaluating the resulting system.. The

structural parameters are given in Table 3.2. These are the same structural parameters as

used in the previous section. Coupling the aerodynamic and structural systems yields the

following continuous time state space system:

−−

−−

=

xxff

qxxff

&

&

&

&&

&

&&

010041.00

000111.082

Computing the eigenvalues over a range of dynamic pressures yields a divergence

stability boundary at a dynamic pressure of 32. The continuous time root locus plot is

given in Figure 3.43. The eigenvalue at zero frequency and at a positive real value shows

that the system becomes unstable through divergence. Aerodynamic eigenvalues are seen

near 38.0 ±−=s .

Figure 3.43 2nd Order Force: Continuous-Time Root Locus

97

Next, this coupled system was converted to a discrete time system. A step size of 0.1 was

used to generate the following discrete time root locus plot.

Figure 3.44 2nd Order Force: Discrete-Time Root Locus

The dynamic pressure for divergence remains at 32. The system identification routine

should show a similar root locus form.

The above continuous system was converted to a discrete time state space form

with use of the matrix exponential. With a discrete step size of 0.1, the discrete state

space form is:

[ ]( )

⋅−⋅−⋅

=−=

=≈=

−−−

=

⋅ ∑

3

431

0

10943.8089.010652.410652.4

784.0715.0089.0963.0

!1

BIAAB

Adtk

eA

dd

n

k

kkdtAd

The total discrete time state space form is:

)()()1( kxBkfAkf ddrrr

+=+

98

Next, this discrete system was implemented in a simple spreadsheet. A simple

chirp excitation signal was input as xr into the system and an output force fr

was

recorded. The input displacement and output force is shown in Figure 3.45. The output

increases in amplitude due to the contribution from x& as the input displacement

frequency increases.

Figure 3.45 2nd Order Force: Dynamic Input and Output

For comparison, a model sensitivity study was performed. The model order was

varied from a 0-2 to a 10-20. The forced RMS error is plotted in Figure 3.46.

99

2 4 6 8 10 12 14 16 18 2010

−8

10−6

10−4

10−2

100

102

0

1

2345

678910

nb Order

RM

S na Order

Figure 3.46 2nd Order Force: RMS Error Study

na orders of zero and one never converge. The first model with an RMS error less than

10-4 is the 3-2 model. Increasing the nb model order can allow for a good model even

with small na model orders. The RMS error converges for all nb orders with na greater

than 3. For this simple system, the RMS error allows for model order determination.

A coupled aero-structural sensitivity study was performed. Figure 3.47 plots the

divergence dynamic pressure versus na and nb model order. This testcase uses the chirp

training signal.

2 4 6 8 10 12 14 16 18 200

5

10

15

20

25

30

35

40

0

1

2 3 45678910

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 3.47 2nd Order Force: Divergence Boundary Study

100

The boundary converges for model orders greater than na equal to 1. The converged

divergence boundary is located at a dynamic pressure of 32.0. This is exactly the

analytical boundary.

An eigenvalue sensitivity study was performed to evaluate the system

identification root locus with respect to the previous analytical result. Figure 3.48 shows

the discrete time eigenvalues for model order of 3-2, 3-5 and 5-10. The eigenvalues

follow the structure given in the analytical case. The system identification and aeroelastic

coupling procedures appears to be working correctly.

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 3.48 2nd Order Force: Root Locus Study

The aerodynamic eigenvalues calculated with system identification appear to “pull”

towards the origin more than the analytical results. The presence of additional

eigenvalues near the unit circle is probably responsible for the difference.

Next, a 20-100 model was evaluated to test the performance of vastly-too-large

models. An eigenvalue plot is shown in Figure 3.49. The important aerodynamic and

101

structural eigenvalues match those of the lower order models and the analytical result.

The spurious eigenvalues that approach the unit circle are discussed in 3.3.11 and 3.3.12.

Figure 3.49 2nd Order Force: Root Locus Study with a Large Model

System identification found the dominant aerodynamics and accurately predicted the

divergence boundary for this simple testcase.

102

CHAPTER 4

4RESULTS

This chapter presents results from aeroservoelastic testcases with the experience

gained from the methodology section.

4.1 Aerodynamic System Identification Training Method

This section briefly outlines the system identification training method.

• Steady State: Run steady euler3d to obtain a steady .unk solution file

• Modeshapes: Run Solids to obtain structural modes. Convert the solids

modeshapes to euler3d .vec file format.

• Convergence: An accurate unsteady solution requires convergence.

Ensure convergence of the unsteady solution with the proper selection of timestep

and iterative parameters in euler3d. Inputting a step response is a good way to test

for unsteady convergence. The following shows the convergence requirements to

ensure solution accuracy. For example, structural motion convergence is ensured

through a sufficient number of structural modes. Adding additional structural

modes is moderately difficult. The “easy” solution convergence requirement

should be checked at every timestep. The CFD grid is difficult to change.

Converged is performed by comparing the solutions to successively refined grids.

All four of these convergence criteria must be satisfied to ensure an accurate

aeroservoelastic solution.

103

Convergence: Difficulty Solution via residuals Easy Time via timestep ... Structure via number of modes ... CFD Grid via grid refinement Hard

• Training Signal Parameters: Select training signal parameters based on

physics and convergence criteria.

• traincreate3d: Create a training directory structure and copies the grid,

modeshape and control files for each mode.

• euler3d: Run unsteady chirp training signal for each mode

• trainassemble3d: Assemble all xn.dat outputs to a single assemble directory.

Rename each xn.dat to case.train#.

• xnmeld: Generate a combined xn.dat file from all case.train# files.

• cfdmdl3d: Generate aerodynamics models

• asemdl3d: Determine forced response and eigenvalues for the coupled model.

4.2 Single Degree of Freedom Divergence

This single DOF divergence testcase investigates the long term aerodynamic

response of an ARMA model. This geometry always exhibits static divergence since the

aerodynamic center of pressure, the point where concentrated lift force acts, is always

ahead of the elastic rotational center for both subsonic and supersonic flows. The steady

state response only involves the dynamic pressure, the spring’s stiffness and the center of

pressure location. An exact solution for this simple geometry is easily derived with

linearized aerodynamics and a steady state force balance. Derivations for subsonic and

supersonic flows are given in Appendix F.

104

Figure 4.1 shows the structural and flow geometry. The airfoil is elastically

constrained at the trailing edge with a spring of strength αK , which provides a restoring

moment proportional to the angle of attack,α . The fluid flows at velocity V with the

elastic rotational center downstream of the airfoil’s leading edge.

α KαV

Figure 4.1 SDOF Divergence Geometry

Implementing this testcase required selecting a sufficiently representative

aerodynamic geometry. Because the testcase needed analysis based on subsonic flows,

the ubiquitous NACA 0012 airfoil was selected with the intention of switching to a

supersonic wedge airfoil if the supersonic flow solution caused difficulties. The CFD

surface tetrahedral grid is illustrated in Figure 4.2.

Figure 4.2 SDOF Divergence CFD Surface Tetrahedral Grid

For unsteady analysis, the CFD grid appears unusually coarse. This coarse grid

should not be a problem in either Mach number regime. First, since the subsonic

105

aerodynamics model will only consider the steady state response, a fine grid is not

needed to capture the high resolution unsteady aerodynamics. Second, for the supersonic

solution, the linearized theoretical flow solution would only need airfoil surface normals.

The structural geometry was developed from simple structural free response

characteristics. This testcase allows for significant freedom in choosing structural

properties because any logical combination of parameters will exhibit static divergence.

Obviously, structural mass and damping only affect the transient structural response, not

the final steady state response. Table 4.1 shows the selected structural parameters. The

undamped natural frequency is 100 s-1 (16 Hertz).

Mass 1/100 slug ft2 Stiffness 100 lbf ft rad-1 Damping 1%

Table 4.1 Single Degree of Freedom Structural Parameters

A generalized displacement of 10 degrees angle of attack was used as one unit of

transpiration deflection. These parameters must be converted to the proper generalized

structural parameters by the transformation described in Appendix E. The resulting

generalized modal parameters are shown in Table 4.2:

Mass 0.0003046

Stiffness 3.046Damping 0.00003046

Table 4.2 Single Degree of Freedom Modal Parameters

Failure to convert between the traditional structural properties and the modal

properties will result in significant structural coupling errors. The following example

shows these errors. Two free vibration time histories, one correct and one incorrect, were

computed for a constant input disturbance. The correct time history uses the transformed

106

modal parameters given in Table 4.2. The incorrect time history uses the traditional

structural parameters given in Table 4.1.

0 20 40 60 80 100 120 140 160 180 20040

20

0

20

40

Time

Displacement

Figure 4.3 SDOF: Correct Displacements

0 20 40 60 80 100 120 140 160 180 2001

0.5

0

0.5

1

Time

Displacement

Figure 4.4 SDOF: Incorrect Displacements

The frequencies are correct, but the magnitudes are not. For multiple degrees of

freedom, the mass and stiffness cross couplings will be corrupted. Correctly identifying

and transforming the coordinate system is important.

The system is ready for aerodynamic analysis at the sub and supersonic Mach

numbers. A Mach number of 2.0 will be used for the supersonic example. The subsonic

example will use a Mach number of 0.6.

4.2.1 Mach 2.0

The supersonic single degree of freedom divergence problem is conducted at

Mach 2.0. The divergence boundary search will be found for two methods: system

identification and linearized supersonic aerodynamics.

The system identification aerodynamics training signal was based on a dc-chirp

training signal. The aerodynamics were calculated from a piston perturbation solution in

euler3d. The piston solution has the advantage of a known aerodynamic input/output

107

relationship. Surface pressure is only a function of the surface normal angle with respect

to the freestream velocity. This aerodynamic system is a zero order system proportional

to, at most, a first derivative of displacement.

Aerodynamics system identification was begun by creating unsteady training

history responses with the euler3d’s piston solver. For comparison, models will be

generated based on the 3211 multistep, the chirp and the dc-chirp training signals.

Training signal parameters and results are summarized in Appendix F. Visual displays of

the displacement, velocity and resulting force are given below for three training signals.

Because the simplified piston solver was used, these simple training signals gave almost

identical results. Using the piston solver as a training signal evaluation tool is not

particularly useful. This “piston” testcase is CFD analog of the zero order force function.

0 50 100 150 200 250 3000

0.05

0.1Displacement

0 50 100 150 200 250 300−1

−0.5

0

0.5

1x 10

−3 velocity

0 50 100 150 200 250 3000

0.5

1x 10

−3 Forces

Figure 4.5 SDOF: Multistep Training Signal

Mach 2.0

0 50 100 150 200 250 300−0.01

−0.005

0

0.005

0.01Displacement

0 50 100 150 200 250 300−2

−1

0

1

2x 10

−3 velocity

0 50 100 150 200 250 3000

1

2

3

4x 10

−4 Forces

Figure 4.6 SDOF: Chirp Training Signal Mach 2.0

0 50 100 150 200 250 3000

0.005

0.01Displacement

0 50 100 150 200 250 300−1

−0.5

0

0.5

1x 10

−3 velocity

0 50 100 150 200 250 3000

1

2

3

4x 10

−4 Forces

Figure 4.7 SDOF: DC-Chirp Training Signal

Mach 2.0

The first set of models is based on the 3211 multistep training signal. Because the

aerodynamics model was trained using a piston solution, the multistep is expected to have

little difficulty with determining an aerodynamics system model. Figure 4.5 clearly

shows the simplistic piston solution. The training signal was processed with

cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to 10-30. Next, asemdl3d was

108

run to determine the force RMS, cross correlation and stability boundary of the models.

The lowest force-RMS occurred with a 4-30 model. The lowest cross correlation

occurred with a 0-5 model. All eigenvalues for a 5-10 model ranging from a dynamic

pressure of 0 to 600 psf are shown in Figure 4.8. The zoomed Figure 4.11 shows the

eigenvalues corresponding to the structural modes. For all models with na less than nb,

static divergence was predicted at 441 psf.

The next model was based on the chirp training signal. The training signal was

processed with cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to 10-30.

Next, asemdl3d was run to determine the force RMS, cross correlation and stability

boundary of the models. An 8-30 model yielded the lowest force-RMS. The lowest cross

correlation occured with a 2-30 model. All eigenvalues for a 5-10 model are shown in

Figure 4.9. The zoomed Figure 4.12 shows the eigenvalues corresponding to the

structural modes. For almost all models, static divergence was predicted at 441 psf.

Models with nb equal to 1 were the exceptions; however, each of these models predicted

divergence within 5 percent of the dominate 441 psf value.

Finally, the dc-chirp training signal was used. The training signal was processed

with cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to 10-30. Next,

asemdl3d was run to determine the force RMS, cross correlation and stability boundary

of the models. The lowest force-RMS occurred with a 9-30 model. The lowest cross

correlation occurred with a 10-3 model. All eigenvalues for a 5-10 model are shown in

Figure 4.10. The zoomed Figure 4.13 shows the eigenvalues corresponding to the

structural modes. Again, static divergence was predicted at 441 psf. Models with nb equal

to 1 were the exception with the worst prediction being a 436 psf with a 10-1 model.

109

Figure 4.8 SDOF: Multistep Eigenvalues

Figure 4.9 SDOF: Chirp Eigenvalues

Figure 4.10 SDOF: DC-Chirp Eigenvalues

Figure 4.11 SDOF: Multistep Zoomed

Eigenvalues

Figure 4.12 SDOF: Chirp Zoomed Eigenvalues

Figure 4.13 SDOF: DC-Chirp Zoomed Eigenvalues

In summary, system identification predicts a static divergence at 441 psf for this

single degree of freedom airfoil at Mach 2.0. It is interesting to note that the three training

signal forms predict nearly identical boundary points. The type of training signal has little

influence on the overall solution accuracy because the piston solution aerodynamics are

so simple. It is noticed that the eigenvalues corresponding to the aerodynamics terms are

somewhat different when plotted over a dynamic pressure range; however, the structural

eigenvalues are identical.

Free response simulations are also available to search for the static divergence

dynamic pressure. The actual dynamic pressure for zero torsional stiffness is not

obtainable because the time history will include an increasing larger static offset

110

displacement as dynamic pressure is increased. This is the classic characteristic obtained

for near-divergent time histories. The effective structural and aerodynamic springs are

balanced at ever increasing angles of attack as the aerodynamic spring becomes stiffer. A

time history plot of displacement is shown in Figure 4.14. Any small increment in

dynamic pressure quickly causes the euler3d implementation of the CFD piston solver to

switch to a non-linear base-pressure routine, which, in this case, effectively eliminates the

divergence phenomenon. The final generalized displacement at 411 psf is 0.5, which

corresponds to 5 degrees angle of attack in the physic model.

0 500 1000 1500 2000 2500 3000 3500 4000 4500 50000

0.5

1Displacement

Figure 4.14 SDOF: Free Response at 411 psf

Now, linearized supersonic airfoil theory is used to determine the theoretical static

divergence dynamic pressure. From linearized supersonic airfoil theory, the center of lift

acts at mid-chord with a lift coefficient 14 2 −= MCl α . As derived in Appendix D,

the dynamic pressure for divergence is:

4110

2

3

−= McK

q α

Applying the structural parameters yields a dynamic pressure for divergence of 433 psf.

Divergence boundaries for the theoretical and the system identification model are

within 3 percent of each other. This close result was not expected because the CFD

aerodynamics grid is a “subsonic” NACA 0012.

111

4.2.2 Mach 0.6

The subsonic divergence boundary prediction for the single degree of freedom

airfoil problem is made at Mach 0.6. The static divergence is predicted using both system

identification and linearized aerodynamics.

System training proceeded similarly to the Mach 2.0 case. Six training signals

were used: the 3211 multistep, the chirp, the dc-chirp, the Schroeder sweep, the strictly

specified Fresnel and the state-space Fresnel. The training signals are shown in the

following group of figures:

112

0 0.5 1 1.5 2 2.5 30

0.05

0.1Displacement

0 0.5 1 1.5 2 2.5 3−0.1

−0.05

0

0.05

0.1velocity

0 0.5 1 1.5 2 2.5 3−2

0

2

4x 10

−3 Forces

Figure 4.15 SDOF: Multistep Training

Signal Mach 0.6

0 0.5 1 1.5 2 2.5 3−0.01

−0.005

0

0.005

0.01Displacement

0 0.5 1 1.5 2 2.5 3−0.2

−0.1

0

0.1

0.2velocity

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−3 Forces

Figure 4.16 SDOF: Chirp Training Signal

Mach 0.6

0 0.5 1 1.5 2 2.5 30

0.005

0.01Displacement

0 0.5 1 1.5 2 2.5 3−0.1

−0.05

0

0.05

0.1velocity

0 0.5 1 1.5 2 2.5 3−2

−1

0

1

2x 10

−3 Forces

Figure 4.17 SDOF: DC-Chirp Training

Signal Mach 0.6

0 0.5 1 1.5 2 2.5 3 3.5 4−0.01

−0.005

0

0.005

0.01Displacement

1

0 0.5 1 1.5 2 2.5 3 3.5 4−0.4

−0.2

0

0.2

0.4velocity

1

0 0.5 1 1.5 2 2.5 3 3.5 4−4

−2

0

2

4x 10

−3 Forces

1

Figure 4.18 SDOF: Schroeder Training Signal Mach 0.6

0 0.5 1 1.5 2 2.5 3 3.50

0.02

0.04

0.06

0.08Displacement

1

0 0.5 1 1.5 2 2.5 3 3.5−0.1

−0.05

0

0.05

0.1Velocity

1

0 0.5 1 1.5 2 2.5 3 3.5−1

0

1

2

3x 10

−3 Forces

1

Figure 4.19 SDOF: Strict Fresnel Mach 0.6

0 0.5 1 1.5 2 2.5 3 3.5 40

0.2

0.4Displacement

1

0 0.5 1 1.5 2 2.5 3 3.5 4−1

−0.5

0

0.5

1Velocity

1

0 0.5 1 1.5 2 2.5 3 3.5 4−5

0

5

10

15x 10

−3 Forces

1

Figure 4.20 SDOF: State Space Fresnel

Mach 0.6

113

The first training signal tested is the 3211 multistep. Figure 4.15 shows the

training signal time history for the multistep. The majority of the output force changes

occur near the step changes in velocity. The training signal was processed with

cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to 10-30. The lowest force-

RMS occured with a 2-25 model. The lowest cross correlation occurred with a 6-9 model.

Model order sensitivity is plotted in Figure 4.21. Dynamic pressure is swept from 0 to

200 psi. Static divergence was predicted in a band from 110 psf to 125 psf. Coupled aero-

structural eigenvalues for the multistep input are plotted in Figure 4.27 in the z plane. The

far right points are the eigenvalues associated with the structural motion, while the others

are aerodynamics. An eigenvalue exists at the Nyquist frequency on the unit circle.

The chirp training signal time history is shown in Figure 4.16. The training signal

was processed with cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to 15-60.

The lowest force-RMS occurred with a 1-33 model. The lowest cross correlation

occurred with a 0-13 model. Eigenvalues for a 5-10 model are shown in Figure 4.28. A

damped eigenvalue exists at the Nyquist frequency. Static divergence was predicted

between a band from 100 psf to 130 psf. Model order sensitivity is plotted in Figure 4.22.

114

0 10 20 30 40 50 600

50

100

150

0

1

2

3

4

5

6

789

1011

1213

1415

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.21 SDOF: Multistep Model Order Sensitivity

0 10 20 30 40 50 600

20

40

60

80

100

120

140

160

180

200

0

1

2

3

456

789101112131415

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.22 SDOF: Chirp Model Order Sensitivity

0 10 20 30 40 50 600

20

40

60

80

100

120

140

160

180

0

1

2

345678

9101112131415

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.23 SDOF: DC-Chirp Model Order Sensitivity

0 10 20 30 40 50 600

20

40

60

80

100

120

140

160

180

0

1

nb Order

Dyn

amic

Pre

ssur

e

na Order

No Stability Boundary Found for na>1

Figure 4.24 SDOF: Schroeder Model Order Sensitivity

0 10 20 30 40 50 600

50

100

150

0

1

2

3

456789101112131415

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.25 SDOF: Strict Fresnel Model Order Sensitivity

0 10 20 30 40 50 600

20

40

60

80

100

120

140

160

180

0

1

2

34

56

7891011121314151617181920

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.26 SDOF: State-Space Fresnel Model Order Sensitivity

The dc-chirp training signal time history is shown in Figure 4.17. The training

signal was processed with cfdmdl3dsplice for model orders of na-nb ranging from 0-1 to

115

15-60. The lowest force-RMS occurred with a 13-19 model. The lowest cross correlation

occurred with a 1-15 model. Eigenvalues for a 5-10 model are shown in Figure 4.29. The

most striking difference between this dc-chirp and the regular chirp is the difference in

expansion rate for the eigenvalues centered on the origin. This non-physical expansion

causes problems with certain testcases at large dynamic pressures. These eigenvalues are

required to exist; however, it is preferred that they remain near the origin. Model order

sensitivity is plotted in Figure 4.22. Static divergence was predicted at 120 psf.

The Schroeder sweep time history is shown in Figure 4.18. For this testcase, the

Schroeder sweep does not provide an accurate representation of the aeroelastic boundary.

Figure 4.24 shows the model order sensitivity of the Schroeder sweep. No stability

boundary was found for nb greater than 1. Additionally, nb convergence does not occur.

A free response simulation indicates static divergence at all dynamic pressures. This

physically impossible result should not occur. The eigenvalues corresponding to the

Schroeder input signal are given in Figure 4.30. An eigenvalue exists on the real axis

outside the unit circle. This is neither desired nor physical. Otherwise, the Schroeder

input signal produces a preferable eigenvalue structure. The Schroeder sweep excitation

appears to be causing a misidentification of the aerodynamics. This problem seems to be

caused by the Schroeder sweep’s parameter sensitivity as discussed in Simon [2000] .

The PSD of the actual input signal used for this testcase revealed gaps in the frequency

spectrum. These frequency subband gaps are known to cause sensitivities [Simon, 2000].

It is also possible that these sensitivities are being introduced by the envelope term used

to start the input signal from zero initial conditions. Unfortunately, a better method for

116

starting the Schroeder sweep without generating non-physical solution transients is not

known.

A strict motion specified Fresnel chirp training time history is given in Figure

4.19. A model order sensitivity study using na from 0 to 15 and nb from 1 to 60 is plotted

in Figure 4.25. Interestingly, the stability boundary prediction appears to converge until

reaching nb of approximately 50. The higher order models are becoming numerically

unstable. The coupled aeroelastic eigenvalues for a 5-10 model are plotted in Figure 4.31

for a dynamic pressure range from 0 to 200 psi. The eigenvalue structure resembles the

dc-chirp’s.

A state-space motion specified Fresnel chirp time history is given in Figure 4.20.

The ending condition for this training signal is a constant velocity, which creates the

linear displacement response after the non-dimensional time of 3.5. A model order

sensitivity study is shown in Figure 4.24. Model order convergence resembles that of the

dc-chirp. Interestingly, the very low nb model orders have significantly more consistent

predictions for this state-space motion specification than for any of the strict motion

specification models. Eigenvalues for the coupled aeroelastic system with a 5-10 model

order are shown in Figure 4.32. An eigenvalue near the Nyquist frequency has appeared.

Also, the near-origin eigenvalue structure is rotated by 45 degrees when compared to the

strict motion specification forms.

117

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.27 SDOF: Multistep

Eigenvalues Mach 0.6

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.28 SDOF: Chirp Eigenvalues

Mach 0.6

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.29 SDOF: DC-Chirp

Eigenvalues Mach 0.6

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.30 SDOF: Schroeder

Eigenvalues Mach 0.6

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.31 SDOF: Strict Fresnel

Eigenvalues Mach 0.6

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Real(z)

Imag

(z)

Figure 4.32 SDOF: State Space Fresnel

Eigenvalues Mach 0.6

118

Linearized airfoil theory is used to determine the theoretical static divergence

dynamic pressure. From linearized subsonic airfoil theory, the center of lift acts at

quarter-chord with a lift coefficient 212 MCl −= πα . The dynamic pressure for

divergence is:

πα

21

320 2

3

McK

q −=

This yields a dynamic pressure for static divergence of 92 psf.

4.3 AGARD 445.6

The AGARD 445.6 testcase presents an aeroelastic structure typical of high

performance aircraft. Aeroelastic characteristics were measured at subsonic and

supersonic Mach numbers. The experiments were conducted in the 1960’s at NASA

Langley [Yates, 1987].

The AGARD structure consists of a cantilevered wing constrained at the root. The

wing has an aspect ratio of 4.0, a quarter chord sweep of 45 degrees and a taper ratio of

0.6. The airfoil is a NACA 65A004. A planform view of the wing with the CFD

tetrahedral mesh is shown in Figure 4.33. The modeshapes and frequencies are shown in

Figure 4.34.

119

Root

Tip

V

Figure 4.33 AGARD: Planform

Mode 1 at 9.6 Hertz

Mode2 at 38.2 Hertz

Figure 4.34 AGARD: Modeshapes

4.3.1 Flutter Boundary

An experimental flutter boundary for the AGARD is reported in [Yates 1987] and

is reproduced in Table 4.3. The actual model description for the AGARD testcase under

consideration is the wall-mounted 2.5 foot span, weakened model #3 tested in air. The

boundary is reported as a flutter speed index instead of dynamic pressure. The

experimental data does not contain any sensitivity or experimental uncertainty

information.

Mach Number

Flutter Speed Index

Dynamic Pressure [psi]

0.499 0.446 0.924 0.678 0.417 0.807 0.900 0.370 0.636 0.960 0.308 0.441 1.072 0.320 0.476 1.141 0.403 0.754

Table 4.3 AGARD 445.6: Experimental Flutter Boundary

The system identification flutter boundary for the AGARD was computed over

the experimental Mach number range with multiple training signals and training

120

methodologies. Corresponding free response simulations evaluated the system

identification predictions for the coupled CFD-structural aeroelastic system.

It is important to remember that the overall objective is to replace the CFD flow

solver with a state space model. Thus, the relevant flutter boundary comparison is

between the CFD and ARMA boundaries not between the ARMA and experimental

boundaries. The objective is for the ARMA model to match the CFD boundary. Because

the structural and aerodynamic grids are not fully converged, it is expected that

differences will occur between the computational and experimental boundaries. Failure to

match experimental data does not indicate failure of the system identification routine.

4.3.1.1 Training Signal Investigation at Mach 0.90

A detailed training signal investigation was performed for the AGARD at Mach

0.90. The flow solution at Mach 0.90 has a shock on the aft outboard portion of the

AGARD wing. The signals investigated are the multistep, chirp, dc-chirp, Schroeder

sweep and the Fresnel chirp.

The multistep sensitivity plot is given in Figure 4.35. The maximum training

modal deflection was 0.5. The system converges with increases in both na and nb orders.

The ARMA predicted flutter boundary is 0.48 psi. The CFD boundary is 0.62 psi. This is

a 23 percent difference error.

121

0 5 10 15 20 25 300

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0

1

23456

7891011121314151617181920

nb Order

na Order

0.636 Experiment

Dynamic Pressure [psi]

Figure 4.35 AGARD: Multistep Sensitivity

The chirp sensitivity plot is given in Figure 4.36. The maximum training

displacement was 0.2. The predictions converge for na>1 and nb>5. The converged

ARMA prediction is 0.59 psi. This is a 5 percent difference error. The stability boundary

prediction decreases after the initial convergence point. This appears to indicate that

better low frequency excitation would be advantageous.

0 5 10 15 20 25 300

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0

1 234567891011121314151617181920

nb Order

Dynamic Pressure [psi]

na Order

0.636 Experiment

Figure 4.36 AGARD: Chirp Sensitivity

122

The dc-chirp sensitivity plot, Figure 4.37, shows improved results. The dc-chirp

was trained with displacements of 0.5. Convergence occurs for nb>5 and na>0 and is

consistent across the entire range of na and nb. Better still, the final dynamic pressure

prediction is 0.62 psi. There is effectively no error when compared to the CFD free

response. An increase in the input signal’s low frequency power seems to have

significantly improved the prediction performance.

0 5 10 15 20 25 300

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0

1 234567891011121314151617181920

nb Order

na Order

0.636 Experiment

Dynamic Pressure [psi]

Figure 4.37 AGARD: DC-Chirp Sensitivity

Increasing the displacement to 4 and 3 units for mode 1 and 2, respectively, yields a

similar, but not quite as converged, flutter boundary. The flow conditions are becoming

nonlinear. The sensitivity plot is given in Figure 4.38. The best prediction appears to be

0.6 psi or approximately 3 percent difference error.

123

0 5 10 15 20 25 300

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0

1

2

345678910

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.38 AGARD: DC-Chirp Large Amplitude Sensitivity

A state-space motion specification excitation with the dc-chirp form is given in Figure

4.39. The forcing function is exactly that used in Figure 4.37 except that the input is

specified through the state-space motion. This was implemented by taking the derivative

of the velocity boundary condition used for the dc-chirp shown in Figure 4.37. The

stability boundary sensitivity with model order is wide. There is no absolute

convergence; however, the predictions are bounded near a flutter boundary of 0.55 psi.

This is an 11 percent difference error.

124

0 5 10 15 20 25 30 35 400

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0

1

2

3

4567

8

910

1112

1314151617

181920

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.39 AGARD: DC-Chirp State Space Methodology Sensitivity

The Schroeder sweep sensitivity plot is given in Figure 4.40. Convergence occurs

but at a slow rate. The best-converged prediction occurs at 0.50 psi or 20 percent

difference error. Overall, the Schroeder gave worse aeroelastic predictions than even the

multistep.

0 5 10 15 20 25 300

0.2

0.4

0.8

1

1.2

1.4

1.6

1.8

0 1

2

3

4

5

67

8

910

111213

14151617181920

nb Order

na Order

0.636 Experiment

Dynamic Pressure [psi]

Figure 4.40 AGARD: Schroeder Sweep Sensitivity

125

The Fresnel chirp sensitivity plot is shown in Figure 4.41. This training signal is

based on a state-space motion specification. The predictions appear to converge to

0.54 psi. This is 13 percent difference error.

0 10 20 30 40 50 600

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0

1

2

3

4567891011121314151617181920

nb Order

Dyn

amic

Pre

ssur

e

na Order

Figure 4.41 AGARD: Fresnel Sensitivity

This training signal sensitivity study shows some interesting results. First, adding

low frequency information to the input signal significantly improves the stability

boundary prediction. This improvement, which is due to additional low frequency power,

was predicted. Second, it is interesting that the state-space motion specification

performed worse than the strict motion specification for the AGARD at Mach 0.90. This

result does not agree with the hypothesis and prediction given in section 3.3.12. The third

interesting result is the importance of smooth higher-order derivative excitation as shown

by comparing the multistep prediction with the other predictions.

4.3.1.2 Mach Number Flutter Boundary

The flutter boundary results across the experimental Mach range are shown in

Figure 4.42. The solid line plots the estimated experimental flutter boundary presented by

126

Yates. The free response boundary is shown as circles, ○. The system identification

boundary is given by stars, *. A free response boundary was not performed at Mach

1.072. For subsonic flow conditions, both the system identification and the free response

solutions are within 10% of the experimental results. Uncertainty analysis was not

presented in Yates, so this excellent prediction fit might be misleading. Still, it must be

stated that the computational predictions are presented without individual tweaking. The

same simulation process, input data files and CFD parameters were used across the entire

Mach range.

0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

Mach Number

Dyn

amic

Pre

ssur

e [p

si]

Figure 4.42 AGARD: Flutter Boundary

The flutter boundary results contain two interesting points. First, the system

identification process accurately predicts the stability boundary when compared to the

CFD solver. Second, the two mode structural assumption appears to become invalid for

supersonic flows. Gordnier [2000] confirms this hypothesis with a mode shape sensitivity

study at Mach 1.141. The flutter boundary decreased as the number of structural modes

increased. Regardless of the comparisons between the computational and experimental

127

flutter boundaries, the ARMA and CFD boundaries match over the range of Mach

numbers.

4.3.2 Sensitivity Studies

Multiple sensitivity studies were performed for the AGARD 445.6 testcase. In

these studies, the discrete time aerodynamics model allowed for quick simulation and

comparison of flow and structural parameters. These studies demonstrate the tremendous

advantages of system models for design and intuition training.

Aerodynamic system models were determined for the AGARD testcase over the

range of Mach numbers reported in the experimental results [Yates, 1987]. At Mach

0.499, a model sensitivity study was performed. The model order was varied from 1-5 to

20-50 and the resulting coupled aerostructural eigenvalues were calculated for increasing

density. The eigenvalues for the first and second modes are plotted in the z plane in

Figure 4.43. The eigenvalue structure converges as the model order increases.

Additionally, the instability occurs in the first mode due to first and second mode

coupling.

Eigenvalues in z-plane at Mach 0.499

Unit Circle, |z|=1

Mode 2

Mode 1

Increasing

Density

Increasing

Density

Figure 4.43 AGARD: Model and Eigenvalue Sensitivity Study Mach 0.499

128

Next, a Mach number sensitivity study was performed. In this study, the effects of

Mach number are investigated for the first mode structural eigenvalues. Figure 4.44 plots

the eigenvalues corresponding to the first structural mode as dynamic pressure increases

at six Mach numbers. The transonic flutter boundary dip seen in the experimental

boundary is explained by a decrease in first mode aerodynamic damping. The

aerodynamic damping increases as Mach number increases beyond Mach 0.96. However,

the eigenvalues stay near the unit circle. For supersonic Mach numbers, structural

damping dominates the location of the stability boundary. From Figure 4.44, the effective

damping follows the unit circle closer as Mach number is increased. At the lower Mach

numbers, a pronounced damping dip occurs before the eigenvalues cross the unit circle

almost perpendicularly. At the higher Mach numbers, the damping dip is less and the

crossing is more acute. This indicates that regardless of the eventual crossing dynamic

pressure, the AGARD at supersonic Mach numbers will always be just marginally stable.

This has sensitivity implications for wind tunnel testing.

1.141 1.072

0.499

0.6780.900

0.960

|z|=1 Figure 4.44 AGARD: Mach Number and Eigenvalue Sensitivity Study

To investigate the sensitivity that structural damping imparts on the AGARD stability

boundary, a sensitivity study was performed at Mach 0.499 and 1.072 for variable

129

structural damping. The eigenvalues for a density sweep were evaluated at damping

ratios of { }%16%8%4%2%1%0=ζ . Figure 4.45 displays the eigenvalues for

the first mode. Figure 4.46 displays the dynamic pressure for the stability boundary

versus damping ratio for the two Mach numbers. The supersonic flow provides little

aerodynamic damping when compared to the subsonic case. Adding structural damping

to the supersonic case appears to have little effect on the frequency.

Mach 0.499 Mach 1.072

Figure 4.45 AGARD: Damping Sensitivity at Mach 0.499 and 1.072

0 2 4 6 8 10 12 14 160.7

0.75

0.8

0.85

0.9

0.95

1

1.05

1.1

Damping Ratio %

Stability Boundary [psi]

Mach 0.499

Mach 1.072

Figure 4.46 AGARD: Damping Sensitivity

A sensitivity study on structural frequency was performed. The free vibration frequency

was modified plus and minus 10% through the modal stiffness while the modal mass

remained constant. This simulates the error occurring due to material inconsistencies and

130

experimental measurement uncertainties. Figure 4.47 displays the dynamic pressure for

the stability boundary versus Mach number and structural frequency.

0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

Mach Number

Dynamic Pressure [psi]+10%

-10%

As-Is

Structural Frequency

Figure 4.47 AGARD: Structural Frequency Sensitivity

Sensitivity studies with the coupled aerostructural systems provide a powerful approach

to understanding complicated aerostructural systems. Additionally, these types of studies

allow for bounding and designing-out experimental uncertainties. Combining a properly

performed experimental study with design information gathered from sensitivity studies

would allow for a significant and useful benchmark aeroelastic testcase.

4.4 Panel Flutter

This panel flutter testcase integrates STARS Solids and the system identification

routine. The geometry is a 2 by 1 foot rectangular aluminum plate 0.10 inches thick

subjected to a Mach 2 flow. STARS Solids generated the modeshapes and frequencies,

which are shown in Figure 4.48. Of the six modes, four are axial and two are

longitudinal.

131

Figure 4.48 Panel: Modeshapes and Frequencies

The euler3d CFD solver uses the finite element grid shown in Figure 4.49. The 2 by 1

plate resides in the finely resolved center of the CFD grid.

Figure 4.49 Panel: CFD Grid

The flow enters the grid on the left hand side of Figure 4.49. To capture the convection,

the downstream portion of the grid is refined.

132

4.4.1 Serial Chirp Training

The traditional system identification process evaluates the system response based

on a continuous series of input signals. All modes are trained during the same CFD run.

The training signal is a frequency-sweep modified chirp. The chirp is based on the

composition of a harmonic function and an envelope function. The sweep rate is specified

by the harmonic function. The highest sweep frequency is based on the envelope

function. Their combination specifies the signal duration.

The panel testcase required six chirps, one for each mode. The chirps are applied

one after the other, as shown in Figure 4.50. For each subplot of displacement, velocity

and force, there are 6 time-history traces, one for each mode.

Figure 4.50 Panel: Serial Training Signal

From the training signal, ARMA models are determined using cfdmdl3d. The

models are verified by forced response simulations, which ensure that the ARMA

predicted forces agree with the actual CFD forces for a 2-7 model. Figure 4.51 shows the

133

actual forces as lines and the ARMA predictions as dots. The predicted forces lie on the

actual forces for all six modes.

The aeroelastic boundaries predicted by asemdl3d are consistent at a density of

0.0163 slug·ft-3. Figure 4.52 shows the resulting coupled aero-structural eigenvalues.

Figure 4.51 Panel: 2-7 Model ARMA Predictions

Figure 4.52 Panel: Eigenvalues

The third mode exits the unit circle first. This result is surprising because traditional 2D

panel flutter theory predicts that dynamic instabilities first occur in modes 1 and 2.

Significant coupling occurs for modes 1 and 2 but the combination does not become

unstable.

For this serial method, the training signal required 290 time units for a solution.

Any disruptions in the training signal, such as a power outage, require the entire signal to

be regenerated from the beginning. For larger problems with more modes, the time

required for a single serial training signal may be weeks or months. This training method

appears to be suboptimal and overly sensitive to possible glitches.

134

4.4.2 Parallel Chirp Training

The parallel system identification process evaluates the system response based on

a separate series of input signals. Each mode is trained during separate CFD runs.

The training signal uses the chirp; however, only one mode is excited per CFD

run. For the plate testcase, six individual runs were made. The six training signals are

plotted in Appendix G.

The six separate training signals are spliced together to form one combined

training signal for the entire testcase. Figure 4.53 shows the six separately numbered

xn.dat files rejoined as one large xn.dat file. The vertical lines indicate the splice points

and the corresponding mode. This file exactly matches the serial training signal.

Figure 4.53 Panel: Parallel Training Signal

As before, ARMA models are fit to the training data. The models and boundary

predictions for this parallel method are identical to the serial method given in the

preceding section. An aeroelastic boundary is found at a density of 0.0163 slug·ft-3. The

135

flutter prediction for the parallel calculation method is exactly the same as the serial

method.

The parallel method allows for a speed increase equal to the number of structural

modes. In this plate test case, each parallel chirp required 47 time units for a total of 282

units of total computation time spread over 6 computers. The serial method requires a

calculation time of 290 units. The timesavings will be even larger for a greater number of

modes. In addition, the training for a given mode may be done at any time and added to

the overall time history as needed without the requirement for running the entire

simulation again.

4.4.3 Free Response Aeroelastic Boundary Validation

The aeroelastic boundary was validated with aero-structure coupled free response

runs. The starting dynamic pressure for this free response validation was determined from

the system identification aeroelastic boundary. Figure 4.54 plots the displacement time

history for a free response with a free stream density of ρ=0.0160 slug·ft-3. This response

is stable.

Figure 4.54 Panel: Free Response at ρ=0.0160 slug·ft-3.

Figure 4.55 plots the displacement time history for a free response with a free stream

density of ρ=0.0163 slug·ft-3. This response is neutrally stable.

136

Figure 4.55 Panel: Free Response at ρ=0.0163 slug·ft-3.

Figure 4.56 plots the displacement time history for a free response with a free stream

density of ρ=0.0166 slug·ft-3. This response is unstable.

Figure 4.56 Panel: Free Response at ρ=0.0166 slug·ft-3.

The free response boundary exactly corresponds to the system-identification

boundary of 0.0163 slug·ft-3. Each free response required approximately 150 units of

time. For the three free responses shown here, the total time was 450 units. Typically, a

search requires significantly more time because an estimated dynamic pressure for flutter

is not available. The traditional free response always will find the stability boundary;

however, the free response search might require significantly more time that a

corresponding system identification search.

4.5 Wing/Flap Control

The general testcase geometry is a two-dimensional wing with a trailing edge control

flap. The entire wing is elastically restrained about an upstream rotation point. The

generalized geometry is shown in Figure 4.57.

137

U

θδ

D eL

Figure 4.57 Wing-Flap: Geometry

4.5.1 Aerodynamic and Structural Representations

The first step with CFD was to create a representation of the flow geometry. A

NACA 0012 airfoil testcase was selected. The CFD grid is given in Figure 4.58.

Significant refinement occurs aft of the airfoil. This was done to capture the

aerodynamics resulting from the shed wake. The grid contains 235-thousand elements.

Figure 4.58 Wing-Flap: CFD Grid

Next, modeshapes corresponding to the wing rotation and the flap control mode

were created with transpiration. The modeshapes are not easily visualized with the CFD

grid. Refer to Figure 4.57 for a schematic view of the modeshapes. Both the wing rotation

and flap deflection modeshapes have a unit deflection of 1 degree.

138

4.5.2 Training

Aerodynamic training data was generated in parallel with the offset dc-chirp. The

chirp frequency swept from zero to 10 percent of Nyquist. Each of the two modes was

individually excited to generate the following two training time histories. Figure 4.59

shows the training time history for the first mode, rotation angle. The figure displays

displacement, velocity and two modal forces. Velocity effects dominate the first mode

force. Second mode forces are almost hidden by the order of magnitude larger first mode

forces; however, the training signal did excite forces in both modes.

0 5 10 15 20 25 30 350

0.5

1

1.5

2Displacement

12

0 5 10 15 20 25 30 35−20

−10

0

10

20Velocity

12

0 5 10 15 20 25 30 35−0.01

−0.005

0

0.005

0.01Forces

12

Figure 4.59 Wing-Flap: Mode 1 Training

0 5 10 15 20 25 30 350

0.5

1

1.5

2Displacement

12

0 5 10 15 20 25 30 35−20

−10

0

10

20Velocity

12

0 5 10 15 20 25 30 35−2

0

2

4x 10

−4 Forces

12

Figure 4.60 Wing-Flap: Mode 2 Training

Training for the second mode is shown in Figure 4.60. Again, the chirp was swept

to 10 percent of Nyquist. The resulting forces due to second mode excitation include

more complex aerodynamics than the first mode. The smaller magnitude symmetrical

output force is the second mode. The flap produces a symmetrical hinge moment

dominated by velocity. Wing moment produced by the flap is clearly dominated by flap

displacement; however, velocity effects are also apparent. A slight perturbation in the

output force occurs between time 5 and 15. This appears to be caused by excitation

pressure waves moving upstream. This bump in the force plot might represent an

139

unsteady pressure distribution effect similar to the Sears problem. Higher frequency

pressure waves appear to be less disruptive.

4.5.3 Controls

The Ricatti method for control gain selection was investigated for this testcase.

System identification and the Ricatti method chosen do not automatically mesh smoothly.

Ricatti assumes state feedback. In the ARMA model, the states contain aerodynamic

forces, which are not available in reality. Those tempted to use a state observer will also

encounter difficulties because the overall ARMA model is controllable, but not

observable. The difficulty lies in not being able to determine the generalized forces from

the output motion. This is analogous to not being able to deduce a unique flow field from

previous motions. The Ricatti method in this form does not appear to be a good choice

for determining control gains with an ARMA model.

In spite of these problems, the Ricatti method was used to find control gains for

the SDOF testcase. This assumes that all states are known so that a solution, regardless of

its applicability, can be found. The Q matrix from the analytical experiment was used;

however, the Ricatti solution requires a positive definite R matrix. This requirement

forced the R matrix to contain limitation data for all inputs, not just the desired flap

deflection angle. To alleviate this problem, the R matrix weighting for non-flap inputs

were set to large magnitudes. Thus, the final 22 by 6 gain matrix only applied inputs into

the flap deflection angle. Figure 4.61 shows the absolute value of the gains for the couple

aero-elastic state vector as determined with the Ricatti solution method. The motion gains

are:

k={2.029, 0.13197, 0.12102, 0.00783}

140

The flap deflection input is:

flapwingflapwinginput δθδθδ && ⋅+⋅+⋅+⋅= 008.0121.0132.003.2

The motion terms are well behaved. Flap angle inputs are dominated by the wing pitch

angle. The control gains for the actual flap angle is an order of magnitude smaller than

wing angle. This is expected and was seen before in the analytical Ricatti solution above.

However, the aerodynamic gains are problematic. First, it would be preferred if these

gains were zero. Second, the gains have large differences in magnitude with oscillating

phase. This indicates an unusually large sensitivity to the aerodynamics.

103

10 2

10 1

100

101

102

103

104

MotionTerms

AerodynamicTerms

--- + +

Gain Magnitudes

Model Coefficient Location Figure 4.61 Wing-Flap: Ricatti Gains

Overall, the Ricatti method for gain determination did not work well with this

ARMA system model. To even find a solution, several questionable modifications were

made to the standard Ricatti approach. Regardless of these modifications, the resulting

gains still contain fundamental errors when coupled with the ARMA system model.

The control gain selection method that did work was the guess-and-test method. The open

loop system was made to dynamically diverge by adjusting dynamic pressure. The open

loop time history response is given in Figure 4.62.

141

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1−40

−20

0

20

40Displacement

1

2

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1−2

−1

0

1x 10

4 Velocity

1

2

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1−50

0

50Forces

1

2

Figure 4.62 Wing-Flap: Open Loop

Now, the control scheme is applied to the system. It required several iterations to

determine and optimize the control gains in order to obtain a stable system. The stable

system resulted from the following gains:

k={3.5, 0, -0.01, -0.04}

A time history response for the above control gains is plotted in Figure 4.63. The control

gains with improved performance included positions and rates of the flap control surface.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8−5

0

5Displacement

12

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8−1000

−500

0

500

1000Velocity

12

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8−2

0

2

4Forces

12

Figure 4.63 Wing-Flap: Closed Loop

The spurious, high frequency eigenvalue problem surfaced with the controlled

system, especially when rates were used in the control gains. The following example

142

shows the same system made stable with the similar control gains; however, the

unwanted eigenvalues eventually cause non-physical dynamic divergence. The gains are:

k={3.5, 0, -0.01, -0.05}.

This gain represents only a small increase in the flap velocity control gain when

compared to the previous control gain. The previous control gain was numerically stable;

this control gain is not stable. The time history and eigenvalues are given in Figure 4.64

and Figure 4.65. Zooming into Figure 4.64 shows that the diverging chatter occurs at the

Nyquist frequency. The control method is working, but the aerodynamic system model

fails to accurately represent reality. Figure 4.65 shows this and other unwanted spurious

eigenvalues distributed around the unit circle. This spurious eigenvalue topic is discussed

in sections 3.3.11 and 3.3.12.

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2−1

−0.5

0

0.5

1Displacement

12

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2−1

−0.5

0

0.5

1x 10

8 Velocity

1

2

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2−2

−1

0

1

2x 10

8 Forces

1

2

Figure 4.64 Wing-Flap: Closed Loop Chatter Time History

−1 −0.5 0 0.5 1−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

Figure 4.65 Wing-Flap: Closed Loop Chatter Eigenvalues

143

CHAPTER 5

5CONCLUSIONS AND RECOMMENDATIONS

5.1 Conclusions

This thesis was concerned with improved system identification for CFD based

aeroservoelastic predictions. The system identification routine was decomposed into

major areas for further study. These areas are system theory, training method and

excitation signals. Improvements were made in all three areas. These improvements

allow less sensitive, faster and more precise predictions. Additionally, a figure of merit

investigation was performed to identify useful model selection routines. Various

aeroelastic configurations were tested to evaluate the improvements and identify further

improvements. The results show that the new system identification techniques give

improved aeroservoelastic predictions.

The system theory section investigated aerodynamic models in relation to linear

systems and the CFD solver. The first important result was the insight that the CFD

solver updates the boundary condition states simultaneously. This has important

implications for excitation signal selection. A set of aerodynamic specific requirements

gives a set of guidelines for system identification models. A canonical form for the

ARMA model was developed.

The training method section investigated the methodology used to generate

system models from raw input training data. The identification data flow is charted. A

144

data redundancy criterion is developed to correlate model size with training data size

requirements. The most important result from this section is the development of a true

parallel training scheme. Parallel training allows for decoupled distributed training of

multiple mode systems, which allows for significant increases in training performance.

The excitation signal section investigated signals for aerodynamic training.

Training signal criteria were formed to assist with selection and development of new

signals. The multistep signals were reviewed and found to have undesirable traits. A new

class of analytic signals based on frequency swept chirps was developed. The high and

low frequency power spectral density of input signal was found to be important. Low

frequency power improves coupled aerostructural prediction accuracy. High frequency

power ensures a physically consistent transfer function, which improves aeroservoelastic

predictions. State-space motion specification did not provide better aeroelastic

predictions when compared to strict motion specification.

Summarizing the results of this thesis shows that improvements were made in

aeroservoelastic prediction quality. This research suggests that further improvements in

linear system models will allow better aerodynamic predictions without resorting to

exotic and less-intuitive non-linear techniques.

5.2 System Identification Recommendations

The results of this thesis advocate the following system identification

methodologies. Use the classical ARMA realization (section 3.1.5). Use parallel training

with the true splice functionality given in cfdmdl3d (section 3.2.7.3). The best excitation

signal is the dc-chirp without an envelope (section 3.3.5). The excitation should remain

harmonic; end the CFD simulation without ending the excitation. Size the excitation

145

parameters based on the Nyquist frequency and the model over-determination factor

(sections 3.2.3 and 3.2.4). If needed, add white noise to the excitation signal after training

to improve the system’s transfer function (section 3.3.11). Evaluate the model fit with the

force prediction RMS (section 3.4.2). Evaluate the effective model quality with the

coupled aeroelastic stability boundary and eigenvalues (section 3.4.4). It is essential to

ensure total simulation convergence of the solution residuals, timestep, structural modes

and CFD grid (section 4.1).

5.3 Recommendations for Further Study

This thesis has investigated aeroservoelastic predictions. Multiple promising areas

of research were found, but not investigated. The following topics are recommended for

further study.

5.3.1 Linear System Theory

The underlying fundamental for aeroelastic predictions is learning more about

linear system theory and applying those theories to unsteady aerodynamics. Generally,

what do we know about unsteady aerodynamics, and how can this be incorporated into

the identification, training and prediction routine?

One guiding assumption has been that the aerodynamics can be considered linear

within certain bounds. This assumption has not been investigated with much vigor, yet it

potentially has significant implications for future developments. How linear are our

typical aerodynamics in the range that we consider? How do CFD unsteady results

compare with the classical and fundamental unsteady developments such as Wagner,

Theodorsen and Sears? Can these classical unsteady results be describabled by linear

146

systems? How can we place a bound on the non-linearities? Is there a way to capture the

nonlinearities as an added on term to the ARMA formulation? How strong are the typical

nonlinearities for geometries and motions of interest? Can a linear system be devised to

ignore weak nonlinearities?

A reoccurring conundrum appears during investigations of boundary conditions

and unsteady aerodynamics. Large step changes in boundary conditions are not

physically possible; however, our discrete time formulation is based on step changes in

boundary conditions. We claim to be unable to model Wagner type problems, yet we also

claim that we are using a linear system. Updating the boundary conditions simultaneously

gives a better representation, yet a question remains. Are unmodeled transients being

introduced by our discrete time choices? How does our discrete time solver compare with

reality with respect to boundary conditions?

The current ARMA formulation is based on equally spaced points. For long-lag

and flows with different wave propagation speeds, are non-uniform delay spacings

worthwhile? Can system models with different update rates be combined? A related

situation occurs for harmonic systems. For example, helicopter rotors contain both high

frequency “blade” aerodynamics and low frequency “rotor” aerodynamics. The two

influence each other, yet they have time constants differing by at least one or two orders

of magnitude. The current system formulation is poorly suited to this type of problem.

A persistent annoyance has been the disparity between the simulation’s sampling

rate and the system dynamics. For solver accuracy and structural integration, both the

CFD and the system model use small timesteps. This corresponds to eigenvalues in the z

plane unit circle behaving almost like the continuous time form. Can we take advantage

147

of the fact that the interesting aeroelastic results occur at comparatively low frequencies?

Would converting to an actual continuous time form be advantageous?

5.3.2 Training Methodology

This thesis also investigated training methodologies including excitation signals.

The following topics appear to be worth further investigation.

The current SVD solver is a stock solution routine. Different methodologies based

on the SVD are claimed to result in better SVD solutions; however, most require

additional statistical information about the SVD data inputs. Is a weighted or correlated

SVD worth the extra effort? Can we use Mahalanobis-distance minimization techniques?

Would normalizing the SVD inputs reduce magnitude scale errors? Would a Total-Least-

Squares or a related methodology assist in more precise model selection?

The current methodology uses a direct approach to determine system coefficients.

It appears that the transfer function directly out of the CFD solution matches the

dominate physics better than the transfer function out of the corresponding ARMA

model. Would forcing the ARMA model to fit a specified transfer function improve the

overall model quality? Could fitting the training data to a transfer function, and then

transforming back, result in less sensitive models? Ljung [1987] and others advocate

frequency domain approaches. Any attempt at this method will encounter the difficulties

caused by frequency domain estimation problems with harmonic windowing.

Input signal sensitivities appear much larger than expected. Small changes in

input signal often cause dramatic changes in model quality. Does this indicate that the

training signals still do not fully excite all of the dynamics? Can the general harmonic

Schroeder problem be solved for arbitrary phasing? How does input signal selection

148

change with the presence of nonlinear physics? Why did state-space motion specification

give poor results? Theory predicted state-space specification would be an improvement

over strict motion specification. Does non-analytic motion specification practically

require more than displacement state information?

Input-output noise was shown to improve model prediction quality. It is suggested

that a study of noise generation with the CFD solver and the ARMA system be

conducted. How much noise do we really have? Can we quantify noise arising from CFD,

model coefficient truncation and machine accuracy? Which type of noise is the most

critical? Can we add noise to the high frequencies without distorting the low frequencies?

5.3.3 Implementation

The following implementation issues appear to be important. An understandably

unglamorous, but important, implementation process is to evaluate, refine and document

the methodologies for usability and ergonomics.

Second order accuracy forcing should be implemented into the aero-structural

integration with the ARMA model. One possibility was developed in the methodology

section, but it has some limitations. The difference in integration scheme is currently the

largest numerical method difference between the CFD and ARMA systems.

The addition of a rigorous controls methodology in euler3d is needed. A simple

example was presented in the results sections; however, real systems contain a multitude

of additional complications, such as control surface actuators and instrumentation

sampling rates. Further investigations into the combination of system identification and

controls are needed.

149

It is recommended that future filenames use a family-name-first approach. All

files for a particular testcase should read and write files using a <testcase>.<type>

structure.It is also recommended that codes and generated data files include a time stamp

and a version number. This will assist in version control and will allow for easier human

communication.

150

BIBLIOGRAPHY

Akaike, H., “A New Look at the Statistical Model Identification,” IEEE Transactions on

Automatic Control, Vol. 19, 1974, pp. 716-723.

Åström, K. J., “On the Choice of Sampling Rates in Parametric Identification of Time

Series,” Information Sciences, Vol. 1, 1969, pp. 273-278.

Bisplinghoff, R. L., Ashley, H., and Halfman, R. L., Aeroelasticity, Dover, 1955.

Björck, Å., Numerical Methods for Least Squares Problems, SIAM, Philadelphia, 1996.

Boeckman, A. A., “Accelerating CFD-Based Aeroelastic Analysis Using Distributed

Processing,” Master’s Thesis, Department of Mechanical and Aerospace

Engineering, Oklahoma State University, 2003.

Boziac, S. M., Digital and Kalman Filtering, Edward Arnold Ltd, London, 1979.

Braun, M. W., Ortiz-Mojica, R. and Rivera, D. E., “Application of Minimum Crest

Factor Multisinusoidal Signals for `Plant-Friendly’ Identification of Nonlinear

Process Systems,” Control Engineering Practice, Vol. 10, 2002, pp. 301-313.

Brenner, M. J., Lind, R. C., and Voracek, D. F., “Overview of Recent Flight Flutter

Testing Research at NASA Dryden,” NASA TM-4792, April 1997.

Brockwell, P. J., and Davis, R. A., Time Series: Theory and Methods, 2nd Edition,

Springer, New York, 1991.

Cai, C. S., and Albrecht, P., “Flutter Derivatives Based Random Parametric Excitation

Aerodynamic Analysis,” Computers and Structures, Vol. 75, 2000, pp. 463-477.

Chen, C. T., Linear System Theory and Design, 3rd Edition, Oxford, New York, 1998.

151

Cowan, T. J., “Efficient Aeroelastic CFD Predictions Using System Identification,”

Masters’s Thesis, Department of Mechanical and Aerospace Engineering,

Oklahoma State University, 1998.

Cowan, T. J., A.S. Arena and K.K. Gupta. “Accelerating Computational Fluid Dynamics

Based Aeroelastic Predictions Using System Identification,” Journal of Aircraft,

Vol. 38, No. 1, Jan-Feb 2001, pp. 81-87.

Cowan, T. J., “Finite Element CFD Analysis of Super-Maneuvering and Spinning

Structures,” Ph.D. Dissertation, Department of Mechanical and Aerospace

Engineering, Oklahoma State Univeristy, 2003.

Deprettere, F. (ed.), SVD and Signal Processing, Elsevier, New York, 1988.

Dowell, E. H. (ed.), et al., A Modern Course in Aeroelasticity, Kluwer Academic

Publishers, Dordrecht, 1995.

Florea, R., Hall, K. C., and Dowell, E. H., “Eigenmode Analysis and Reduced-Order

Modeling of Unsteady Transonic Potential Flow Around Airfoils,” Journal of

Aircraft, Vol. 37, No. 3, 2000, pp. 454-462.

Gevers, M. R., “ARMA models, their Kronecker indices and their McMillan degree,”

International Journal of Control, Vol. 43, No. 6, 1986, pp. 1745-1761.

Gordnier, R. E., Melville, R. B., “Transonic Flutter Simulations Using an Implicit

Aeroelastic Solver,” Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 872-879.

Gupta, K. K., “STARS - An Integrated, Multidisciplinary, Finite-Element, Structural,

Fluids, Aeroelastic, and Aeroservoelastic Analysis Computer Program,” NASA

TM-4795, April 2001.

152

Irani, M., and Anandan, P. “Factorization with Uncertainty,” International Journal of

Computer Vision, Vol. 49, No. 2-3, 2002, pp. 101-116.

Hamel, P. G., and Jategaonkar, R. V., “Evolution of Flight Vehicle System

Identification,” Journal of Aircraft, Vol. 33, No. 1, 1996, pp. 9-28.

Hunter, J. P., “An Efficient Method for Time-Marching Supersonic Flutter Predictions

Using CFD,” Master’s Thesis, Department of Mechanical and Aerospace

Engineering, Oklahoma State University, 1997.

Ljung, L., System Identification: Theory for the User, Prentice Hall, Upper Saddle River,

1987.

McCormack, A. S., and Godfrey, K. R., and Flower, J. O., “Design of Multilevel

Multiharmonic Signals for System Identification,” IEEE Proceedings on Control

Theory and Applications, Vol. 142, No. 3, May 1995, pp. 247-252.

Mäkilä, P. M., “On Linear Models for Nonlinear Systems,” Automatica, Vol. 39, Issue. 1,

Jan. 2003, pp. 1-13.

Mehra, R. K., “Optimal Inputs for Parameter Estimation in Dynamic Systems - Survey

and New Results,” IEEE Transactions on Automatic Control, Vol. 19, No. 6, Dec.

1974, pp. 753-768.

Mehra, R. K., “Optimal Inputs for Linear System Identification,” IEEE Transactions on

Automatic Control, Vol. 19, No. 3, June 1974, pp. 192-200.

Miller, D. F., “A Time-Domain Approach to z-Domain Model Identification,”

International Journal of Control, Vol. 44, No. 5, Nov 1986, pp. 1285-1295.

Ninness, B., and Goodwin, G., “Estimation of Model Quality,” Automatica, Vol. 31,

Issue. 12, Dec. 1995, pp. 1771-1797.

153

Parameswaren, V., and Baeder, J. D., “Indicial Aerodynamics in Compressible Flow –

Direct Computational Fluid Dynamic Calculations,” Journal of Aircraft, Vol. 34,

No. 1, Jan-Feb 1997, pp. 131-133.

Press, W. H., Teukolsky, S. A., Vettering, W. T., and Flannery, B. P., Numerical recipes

in Fortran 77, 2nd edition, Vol. 1, Cambridge University Press, 1992.

Reisenthel, P. H., “Development Of A Nonlinear Indicial Model Using Response

Functions Generated By A Neural Network,” AIAA-97-0337, 1997.

Schoukens, J., Pintelon, R., Van Der Ouderaa, E., and Renneboog, J., “Survey of

Excitation Signals for FFT Based Signal Analyzers,” IEEE Transactions on

Instrumentation and Measurement, Vol. 37, No. 2, 1988, pp. 342-352.

Schroeder, M. R., “Synthesis of Low-Peak-Factor Signals and Binary Sequences with

Low Autocorrelation,” IEEE Transactions on Information Theory, Vol. 16, Jan

1970, pp. 85-89.

Simon, G., and Schoukens, J., “Robust Broadband Periodic Excitation Design,” IEEE

Transactions on Instrumentation and Measurement, Vol. 49, No. 2, April 2000,

pp. 270-274.

Van Der Ouderaa, E., Schoukens, J., and Renneboog, J., “Peak Factor Minimization of

Input and Output Signals of Linear Systems,” IEEE Transactions on

Instrumentation and Measurement, Vol. 37, No. 2, 1988, pp. 207-212.

Worden, K., “On the Over-Sampling of Data for System Identification,” Mechanical

Systems and Signal Processing, Vol. 9, No. 3, 1995, pp. 287-297.

Yates, E. C., “AGARD Standard Aeroelastic Configurations for Dynamic Response.

Candidate Configuration I.-Wing 445.6,” NASA TM-100492, August 1987.

154

Young, P., and Patton, R. J., ”Comparison of Test Signals for Aircraft Frequency Domain

Identification,” Journal of Guidance, Control, and Dynamics, Vol. 13, No. 3,

May-June 1990, pp. 430-438.

Zwillinger, D., CRC Standard Mathematical Tables and Formulae, CRC Press, Boca

Raton, 1996.

155

APPENDIX A:

ARMA MODEL .MDL STRUCTURE

This section describes the ARMA model structure as stored in the .mdl file.

$ Mach # tsamp rhoinf { }∞∆ ρ,, tM $ offsets { }nrFFF ,,, 21 L $ na nb nr { }nrnbna ,, $ Model parameters

)(

)(

)1(

)1(

1

1

natf

natf

tf

tf

nr

nr

M

M

M

Number of Coefficients: nrna ⋅

)(

)(

)1(

)(

1

11

natf

natf

tf

tq

nr

nr

M

M

M

Number of Coefficients: 2nrnb ⋅

156

APPENDIX B:

STARS IMPLEMENTATION

cfdmdl3d: cfdmdl3d is the system identification generation routine that converts

between the raw CFD training response and the ARMA model files. The basic code was

supplied from an euler3d port of the original STARS mg2 system identification code

from Tim Cowan’s Master’s thesis [1998]. The following changes were made.

The code was rewritten and re-commented to euler3d specifications where necessary. The

SVD routine was validated with matlab for simple and complex input files. The detrend

routine was changed to allow for state-space motion specification by adding a small

perturbation limitation rather than a strict non-zero starting condition. The entire

modelgen routine was rewritten for easier visualization of the ARMA model structure

data. Functionality for multiple parallel training signals was implemented. A true-splice

functionality was implemented. For debugging, visualization and performance, matrix

operations are now performed as subroutines rather than inline coding.

Input: <case>.train<#>

<case>.con

Output: <case>.mdl<#>

157

Strict Motion Specification Excitation Signals: New strict motion specification

exciation signals were added to euler3d. Subroutines were added to the asedrv.f code to

allow for motion specification through the vector file, <case>.vec. Additional sweep

parameters were added to the euler3d control file.

Excitation Signals:

Serial Chirp IBXN 5

Parallel Chirp IBXN 6

Parallel DC-Chirp IBXN 7

Schroeder Sweep IBXN 8

Fresnel Chirp IBXN 9

Sweep Parameters:

omega linear sweep rate

ratio sweep length

displ amplitude

158

asemdl3d: asemdl3d is the code that allows for aerodynamic and aeroelastic

computations based on the aerodynamic model. The routine accepts both single and

multiple aerodynamic models. The routine allows for: forced motion response to arbitrary

modal motion as read from an xn.dat file; aero-structural free response calculations that

mimic the euler3d CFD solver with the aerodynamic model. The routine also allows for

coupled aeroelastic eigenvalue sweeps over a dynamic pressure range.

Input Files:

<case>.mdl<#>

<case>.con

<case>.vec

<xn.dat>

Output:

<case>.eig aero-structural eigenvalues

<case>out.dat forced response output

<case>.RMS model performance output

<case>.amtx model aerodynamic state transition matrices

<case>.smtx structural state transition matrices

sensitivity.dat stability boundaries for multiple models

159

xnmeld3d: xnmeld3d combines multiple parallel training signals into one single

signal. One comparison file is desired for parallel identification performance evaluation.

Input Files:

<case>.train<#>

Output File:

<xn.dat>

trainassemble3d: trainassemble3d retrieves the excitation data from each parallel

training directory and assembles training data into a single directory. Individual xn.dat

files are renamed to the appropriate <case>.train<#> file.

Input Files:

.\mode<#>\xn.dat

Output Files:

.\assemble\<case>.train<#>

160

APPENDIX C:

FREQUENCY SWEEP PARAMETER SELECTION

Proper training with the frequency sweep training signals requires proper

parameter selections. The following shows a general guideline for selecting the frequency

parameters for a linear frequency sweep signal.

The total number of training datapoints determines the overdetermined system

factor. More data allows better statistical fits at the expense of training length. Typically,

500 to 1000% overdetermination yields converged system models. The number of

datapoints required is:

( )

⋅⋅+⋅=

100%inedOverdetermdatapoints 2 nanrnanr

For accurate aero-structural integration, each cycle of the highest frequency must

have a sufficient number of points per cycle. The selected timestep, dt, must be at least

smaller than the following:

( )cycleper pts11

min

⋅≤∗

freqdt

For physical flow solution convergenc, the frequency sweep should not exceed a certain

percentage of the Nyquist frequency based on the CFD timestep. Typically, one-tenth of

Nyquist is sufficient.

( )tioequency RaNyquist Frdt

⋅=

= ∗

NyquistSweep

Nyquist2

21

ωω

πω

161

From the above timestep and datapoint calculations, the non-dimensional time required

for the training signal is:

( ) ∗⋅= dtnr

sdata pointratio

This parameter is set into the euler3d control file.

The frequency sweep range parameter, omega, determines the increase rate in

frequency:

Sweep1

21 ω⋅⋅=

ratioomega

This parameter is set into the euler3d control file.

For parallel training, the training length for each mode is:

( )points settlingsolutionnrpoints datanstp +=

The nstp parameter is set into the euler3d control file. Solution settling time exists to

assist with capturing the aerodynamic terms without motion influences. Typically, the

solution should settle to “almost” steady state.

The training signal’s amplitude is selected based on expected linearity limitations.

The magnitude should be sufficient to excite the dominant linear aerodynamics without

creating strong nonlinear responses. Naturally, the unit modeshape “deflection”

determines the overall magnitude. There is no general guideline for amplitude.

162

APPENDIX D:

1D DIVERGENCE DERIVATIONS

Mach 2.0 Moment about trailing edge

ααKMce

eLM

==

⋅=

21

Lift at AC

qM

cL

qCcwL l

⋅−

⋅=

⋅⋅⋅=

14

2

251 α

Summation of Moments

( ) αααKcq

Mc =⋅

−⋅ 2

12

251

14

Solve for q

4110

2

3

−= McK

q α

Enter geometry

psfq 433=

163

Mach 0.6 Moment about trailing edge

ααKMce

eLM

==

⋅=

43

Lift at AC

qCcwL l ⋅⋅⋅=

( ) απααKcq

Mc =⋅

−⋅ 4

32

251

12

Solve for q

πα

21

320 2

3

McK

q −=

164

APPENDIX E:

STRUCTURAL MODE CONVERSION

This section explains how modal mass and stiffness values for Euler3d and

STARS are determined from actual structural values. The primary problem in calculating

these values consists of differences between the mode shape and reality. The fundamental

structural equation of motion for a two dimensional problem is:

=

+

αααα

α

αα ffh

K00Kh

ISSm hh

The conversion between the generalized displacement q and the actual motion is a

linear axis transformation. A similar transformation is made for generalized forces:

=

2

1

180 qq

00zhλα π

=

2

1180

z1

h

GG

00

ff

πλα

Substitute, rearrange and simplify to obtain the relevant modal parameters:

=

+

2

1180

z1

2

1

180

h

2

1

180 GG

00

qq

00z

K00K

qq

00z

ISSm

πλπ

απ

αα

α

λλ

( ) ( )

=

⋅⋅

+

⋅⋅⋅⋅⋅⋅

2

1

2

122

180

2h

2

122

180180

1802

GG

qq

K00zK

qq

IzSzSzm

λλλλ

πα

πα

πα

πα

The modal parameters are now compatible with STARS and Euler3d.

165

APPENDIX F:

SINGLE DEGREE OF FREEDOM DIVERGENCE

Configuration Files Mach 0.6 DC-Chirp Mach 2.0 DC-Chirp &control dt = 0.01, gamma = 1.40d0, diss = 1.00d0, cfl = 0.5d0, mach = 0.60d0, alpha = 0.0d0, beta = 0.0d0, refdim = 1.0d0, nstp = 350, nout = 10, ncyc = 60, isol = 2, idiss = 1, ipnt = 1, isize = 50, omega = 6.042, ratio = 3, displ = 0.01, istrt = .true., iaero = .true., idynm = .false., ielast = .true., ifree = .true., iforce = .false., nr = 1, ainf = 1100, rhoinf = 2.40E-3, =? /

&control dt = 1.0, gamma = 1.40d0, diss = 1.00d0, cfl = 0.5d0, mach = 2.00d0, alpha = 0.0d0, beta = 0.0d0, refdim = 1.0d0, nstp = 400, nout = 100, ncyc = 60, isol = 3, idiss = 1, ipnt = 1, isize = 50, omega = 6.042E-4, ratio = 260, displ = 0.01, istrt = .false., iaero = .true., idynm = .false., ielast = .true., ifree = .true., iforce = .false., nr = 1, ainf = 1100, rhoinf = 1.80E-7, =? /

166

Modeshape Vector File $ Number of elastic modes (nr) 1 $ Mass matrix for elastic modes (nr x nr) 3.046174E-4 $ Damping matrix for elastic modes (nr x nr) 3.0E-6 $ Stiffness matrix for elastic modes (nr x nr) 3.046174 $ ICs for elastic modes (x1....xn, vx1...vxn) 0.0d0 0.0d0 $ IBXN for elastic modes (nr) 7 $ Elastic modes vectors (nwl 2) x nr 0.1519224903E-01 0.0000000000E+00 0.1736482010E+00 0.1519224903E-01 0.0000000000E+00 0.1736482010E+00 0.5317286142E-01 0.0000000000E+00 0.6077685871E+00 -0.1291340994E+00 0.0000000000E+00 -0.1476009510E+01

167

APPENDIX G:

PANEL FLUTTER

Plate Parallel Chirp Training Responses at Mach 2.0

Mode 1

Mode 2

Mode 3

Mode 4

Mode 5

Mode 6