+ All Categories
Home > Documents > Stability Control EtkinReid

Stability Control EtkinReid

Date post: 07-Oct-2014
Category:
Upload: amber-wright
View: 323 times
Download: 2 times
Share this document with a friend
390
j, Dynamics of Flight Stability and Control
Transcript
Page 1: Stability Control EtkinReid

•j,

Dynamics of FlightStability and Control

Page 2: Stability Control EtkinReid

ACQUIsmONS EDITOR Cliff RobichaudASSISTANT EDITOR Catherine BeckhamSENIOR PRODUCTION EDITOR Cathy RondaCOVER DESIGNER Lynn Rogan

MANUFACTURING MANAGERILLUSTRATION COORDINATOR

Susan StetzerGene Aiello

This book was set in Times Roman by General Graphic Services, and printed and bound byHamilton Printing Company. The cover was printed by Hamilton Printing Company.

Recognizing the importance of preserving what has been written, it is a policy of John Wiley &Sons, Inc. to have books of enduring value published in the United States printed on acid-freepaper, and we exert our best efforts to that end.

The paper in this book was manufactured by a mill whose forest management programs includesustained yield harvesting of its timberlands. Sustained yield harvesting principles ensure that thenumber of trees cut each year does not exceed the amount of new growth.

Copyright © 1996, by John Wiley & Sons, Inc.

All rights reserved. Published simultaneously in Canada.

No part of this publication may be reproduced, stored in a retrieval system or transmittedin any form or by any means, electronic, mechanical, photocopying, recording, scanningor otherwise, except as permitted under Sections 107 or 108 of the 1976 United StatesCopyright Act, without either the prior written permission of the Publisher, orauthorization through payment of the appropriate per-copy fee to the CopyrightClearance Center, 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400, fax(978) 646-8600. Requests to the Publisher for permission should be addressed to thePermissions Department, John Wiley & Sons, Inc., 111River Street, Hoboken, NJ 07030,(201) 748-6011, fax (201) 748-6008.

To order books or for customer service please, call1(800)-CALL-WILEY (225-5945).

library of Congress Cataloging-in-Publication Data

Etkin, Bernard.Dynamics of flight: stability and control I Bernard Etkin, Lloyd

Duff Reid.-3rd ed,p, em.

Includes bibliographical references (p. ).ISBN 978-0-471-03418-6 (cloth: alk. paper)1. Aerodynamics. 2. Stability of airplanes. 1. Reid, Lloyd D.

II. Title.

TL570.E751995629.132'3-dc20 95-20395

CIP

Printed in the United States of America

10

Page 3: Stability Control EtkinReid

To the men and women of science andengineering whose contributions to aviationhave made it a dominant force in shaping thedestiny of mankind, and who, with sensitivity

and concern, develop and apply theirtechnological arts toward bettering the future.

Page 4: Stability Control EtkinReid
Page 5: Stability Control EtkinReid

PREFACE

The first edition of this book appeared in 1959-indeed before most students readingthis were born. It was well received both by students and practicing aeronautical en-gineers of that era. The pace of development in aerospace engineering during thedecade that followed was extremely rapid, and this was reflected in the subject offlight mechanics. The fust author therefore saw the need at the time for a more ad-vanced treatment of the subject that included the reality of the round rotating Earthand the real unsteady atmosphere, and hypersonic flight, and that reflected the explo-sive growth in computing power that was then taking place (and has not yet ended!).The result was the 1972 volume entitled Dynamics of Atmospheric Flight. That treat-ment made no concessions to the needs of undergraduate students, but attemptedrather to portray the state of the art of flight mechanics as it was then. To meet theneeds of students, a second edition of the 1959 book was later published in 1982. It isthat volume that we have revised in the present edition, although in a number of de-tails we have preferred the 1972 treatment, and used it instead.

We have retained the same philosophy as in the two preceding editions. That is,we have emphasized basic principles rooted in the physics of flight, essential analyti-cal techniques, and typical stability and control realities. We continue to believe, asstated in the preface to the 1959 edition, that this is the preparation that students needto become aeronautical engineers who can face new and challenging situations withconfidence.

This edition improves on its predecessors in several ways. It uses a real jet trans-port (the Boeing 747) for many numerical examples and includes exercises for stu-dents to work in most chapters. We learned from a survey of teachers of this subjectthat the latter was a sine qua non. Working out these exercises is an important part ofacquiring skill in the subject. Moreover, some details in the theoretical developmenthave been moved to the exercises, and it is good practice in analysis for the studentsto do these.

Students taking a course in this subject are assumed to have a good backgroundin mathematics, mechanics, and aerodynamics, typical of a modem university coursein aeronautical or aerospace engineering. Consequently, most of this basic materialhas been moved to appendices so as not to interrupt the flow of the text.

The content of Chapters 1 through 3 is very similar to that of the previous edi-tion. Chapter 4, however, dealing with the equations of motion, contains two verysignificant changes. We have not presented the nondimensional equations of motion,but have left them in dimensional form to conform with current practice, and we haveexpressed the equations in the state vector form now commonly used. Chapter 5, onstability derivatives, is almost unchanged from the second edition, and Chapters 6and 7 dealing with stability and open loop response, respectively, differ from theirpredecessors mainly in the use of the B747 as example and in the use of the dimen-sional equations. Chapter 8, on the other hand, on closed loop control, is very muchexpanded and almost entirely new. This is consistent with the much enhanced impor-tance of automatic flight control systems in modem airplanes. We believe that the

vii

Page 6: Stability Control EtkinReid

viii Preface

student who works through this chapter and does the exercises will have a good graspof the basics of this subject.

The appendices of aerodynamic data have been retained as useful material forteachers and students. The same caveats apply as formerly, The data are not intendedfor design, but only to illustrate orders of magnitude and trends. They are provided togive students and teachers ready access to some data to use in problems and projects.

We acknowledge with thanks the assistance of our colleague, Dr. 1. H. de Leeuw,who reviewed the manuscript of Chapter 8 and made a number of helpful sugges-tions.

On a personal note-as the first author is now in the 11th year of his retirement,this work would not have been undertaken had Lloyd Reid not agreed to collaboratein the task, and if Maya Etkin had not encouraged her husband to take it on and sup-ported him in carrying it out.

In tum, the second author, having used the 1959 edition as a student (with thefirst author as supervisor), the 1972 text as a researcher, and the 1982 text as ateacher, was both pleased and honored to work with Bernard Etkin in producing thismost recent version of the book.

TorontoDecember, 1994

Bernard EtkinLloyd Duff Reid

Page 7: Stability Control EtkinReid

CONTENTS

CHAPTER 1 Introduction1.1 The Subject Matter of Dynamics of Flight1.2 The Tools of Flight Dynamicists 51.3 Stability, Control, and Equilibrium 61.4 The Human Pilot 81.5 Handling Qualities Requirements 111.6 Axes and Notation 15

CHAPTER 2 Static Stability and Control-Part 12.1 General Remarks 182.2 Synthesis of Lift and Pitching Moment 232.3 Total Pitching Moment and Neutral Point 292.4 Longitudinal Control 332.5 The Control Hinge Moment 412.6 Influence of a Free Elevator on Lift and Moment 442.7 The Use of Tabs 472.8 Control Force to Trim 482.9 Control Force Gradient 512.10 Exercises 522.11 Additional Symbols Introduced in Chapter 2 57

CHAPTER 3 Static Stability and Control-Part 23.1 Maneuverability-Elevator Angle per g 603.2 Control Force per g 633.3 Influence of High-Lift Devices on Trim and Pitch Stiffness 643.4 Influence of the Propulsive System on Trim and Pitch Stiffness 663.5 Effect of Structural Flexibility 723.6 Ground Effect 743.7 CG Limits 743.8 Lateral Aerodynamics 763.9 Weathercock Stability (Yaw Stiffness) 773.10 Yaw Control 803.11 Roll Stiffness 813.12 The Derivative C1/3 833.13 Roll Control 863.14 Exercises 893.15 Additional Symbols Introduced in Chapter 3 91

CHAPTER 4 General Equations of Unsteady Motion4.1 General Remarks 934.2 The Rigid-Body Equations 93

ix

Page 8: Stability Control EtkinReid

x Contents

4.3 Evaluation of the Angular Momentum h 964.4 Orientation and Position of the Airplane 984.5 Euler's Equations of Motion 1004.6 Effect of Spinning Rotors on the Euler Equations 1034.7 The Equations Collected 1034.8 Discussion of the Equations 1044.9 The Small-Disturbance Theory 1074.10 The Nondimensional System 1154.11 Dimensional Stability Derivatives 1184.12 Elastic Degrees of Freedom 1204.13 Exercises 1264.14 Additional Symbols Introduced in Chapter 4 127

CHAPTER 5 The Stability Derivatives5.1 General Remarks 1295.2 The a Derivatives 1295.3 The u Derivatives 1315.4 The q Derivatives 1355.5 The a Derivatives 1415.6 The ,B Derivatives 1485.7 The p Derivatives 1495.8 The r Derivatives 1535.9 Summary of the Formulas 1545.10 Aeroelastic Derivatives 1565.11 Exercises 1595.12 Additional Symbols Introduced in Chapter 5 160

CHAPTER 6 Stability of Uncontrolled Motion6.1 Form of Solution of Small-Disturbance Equations 1616.2 Longitudinal Modes of a Jet Transport 1656.3 Approximate Equations for the Longitudinal Modes 1716.4 General Theory of Static Longitudinal Stability 1756.5 Effect of Flight Condition on the Longitudinal Modes of a Subsonic Jet

Transport 1776.6 Longitudinal Characteristics of a STOL Airplane 1846.7 Lateral Modes of a Jet Transport 1876.8 Approximate Equations for the Lateral Modes 1936.9 Effects of Wind 1966.10 Exercises 2016.11 Additional Symbols Introduced in Chapter 6 203

CHAPTER 7 Response to Actuation of the Controls-Open Loop7.1 General Remarks 2047.2 Response of LinearlInvariant Systems 2077.3 Impulse Response 210

Page 9: Stability Control EtkinReid

Contents xi

7.4 Step-Function Response 2137.5 Frequency Response 2147.6 Longitudinal Response 2287.7 Responses to Elevator and Throttle 2297.8 Lateral Steady States 2377.9 Lateral Frequency Response 2437.10 Approximate Lateral Transfer Functions 2477.11 Transient Response to Aileron and Rudder 2527.12 Inertial Coupling in Rapid Maneuvers 2567.13 Exercises 2567.14 Additional Symbols Introduced in Chapter 7 258

CHAPTER 8 Closed-Loop Control8.1 General Remarks 2598.2 Stability of Closed Loop Systems 2648.3 Phugoid Suppression: Pitch Attitude Controller 2668.4 Speed Controller 2708.5 Altitude and Glide Path Control 2758.6 Lateral Control 2808.7 Yaw Damper 2878.8 Roll Controller 2908.9 Gust Alleviation 2958.10 Exercises 3008.11 Additional Symbols Introduced in Chapter 8 301

APPENDIX A Analytical ToolsA.1 Linear Algebra 303A.2 The Laplace Transform 304A3 The Convolution Integral 309A4 Coordinate Transformations 310A5 Computation of Eigenvalues and Eigenvectors 315A6 Velocity and Acceleration in an Arbitrarily Moving Frame 316

APPENDIX B Data for Estimating Aerodynamic Derivatives 319

APPENDIX C Mean Aerodynamic Chord, Mean Aerodynamic Center, andc.; 357

APPENDIX D The Standard Atmosphere and Other Data 364

APPENDIX E Data For the Boeing 747-]00 369

References 372

Index 377

Page 10: Stability Control EtkinReid
Page 11: Stability Control EtkinReid

CHAPTER 1

Introduction

1.1 The Subject Matter of Dynamics of FlightThis book is about the motion of vehicles that fly in the atmosphere. As such it be-longs to the branch of engineering science called applied mechanics. The three itali-cized words above warrant further discussion. To begin with fly-the dictionary defi-nition is not very restrictive, although it implies motion through the air, the earliestapplication being of course to birds. However, we also say "a stone flies" or "an ar-row flies," so the notion of sustention (lift) is not necessarily implied. Even the at-mospheric medium is lost in "the flight of angels." We propose as a logical scientificdefinition that flying be defined as motion through a fluid medium or empty space.Thus a satellite "flies" through space and a submarine "flies" through the water. Notethat a dirigible in the air and a submarine in the water are the same from a mechani-cal standpoint-the weight in each instance is balanced by buoyancy. They are sim-ply separated by three orders of magnitude in density. By vehicle is meant any flyingobject that is made up of an arbitrary system of deformable bodies that are somehowjoined together. To illustrate with some examples: (1) A rifle bullet is the simplestkind, which can be thought of as a single ideally rigid body. (2) A jet transport is amore complicated vehicle, comprising a main elastic body (the airframe and all theparts attached to it), rotating subsystems (the jet engines), articulated subsystems (theaerodynamic controls) and fluid subsystems (fuel in tanks). (3) An astronaut attachedto his orbiting spacecraft by a long flexible cable is a further complex example of thisgeneral kind of system. Note that by the above definition a vehicle does not necessar-ily have to carry goods or passengers, although it usually does. The logic of the defi-nitions is simply that the underlying engineering science is common to all these ex-amples, and the methods of formulating and solving problems concerning the motionare fundamentally the same.

As is usual with definitions, we can fmd examples that don't fit very well. Thereare special cases of motion at an interface which we mayor may not include in fly-ing-for example, ships, hydrofoil craft and air-cushion vehicles (ACY's). In thisconnection it is worth noting that developments of hydrofoils and ACY's are fre-quently associated with the Aerospace industry. The main difference between thesecases, and those of "true" flight, is that the latter is essentially three-dimensional,whereas the interface vehicles mentioned (as well as cars, trains, etc.) move approxi-mately in a two-dimensional field. The underlying principles and methods are stillthe same however, with certain modifications in detail being needed to treat these"surface" vehicles.

Now having defined vehicles and flying, we go on to look more carefully at whatwe mean by motion. It is convenient to subdivide it into several parts:

1

Page 12: Stability Control EtkinReid

2 Chapter 1. Introduction

Aerodynamics

Mechanics of

\ r--+ Vehiclerigid bodies design

Mechanics of~+~ FLIGHT Vehicle

elastic structures DYNAMICS operation

LIHuman pilot L......+ Pilotdynamics training

Applied mathematics,machine computation

J,Performance Stability and Aeroelasticity Navigation and(trajectory, control (handling (control. structural

maneuverability) qualities. airloads) integrity) guidance

Figure 1.1 Block diagram of disciplines.

Gross Motion:1. Trajectory of the vehicle mass center. 1

2. ''Attitude'' motion, or rotations of the vehicle "as a whole."

Fine Motion:3. Relative motion of rotating or articulated subsystems, such as engines, gyro-

scopes, or aerodynamic control surfaces.4. Distortional motion of deformable structures, such as wing bending and twist-

ing.5. Liquid sloshing.

This subdivision is helpful both from the standpoint of the technical problems as-sociated with the different motions, and of the formulation of their analysis. It issurely self-evident that studies of these motions must be central to the design and op-eration of aircraft, spacecraft, rockets, missiles, etc. To be able to formulate and solvethe relevant problems, we must draw on several basic disciplines from engineeringscience. The relationships are shown on Fig. 1.1. It is quite evident from this figurethat the practicing flight dynamicist requires intensive training in several branches ofengineering science, and a broad outlook insofar as the practical ramifications of hiswork are concerned.

In the classes of vehicles, in the types of motions, and in the medium of flight,this book treats a very restricted set of all possible cases. It deals only with the flight

lIt is assumed that gravity is uniform, and hence that the mass center and center of gravity (CG) arethe same point.

Page 13: Stability Control EtkinReid

1.1 The Subject Matter of Dynamics of Flight 3

of airplanes in the atmosphere. The general equations derived, and the methods of so-lution presented, are however readily modified and extended to treat many of theother situations that are embraced by the general problem.

All the fundamental science and mathematics needed to develop this subject ex-isted in the literature by the time the Wright brothers flew. Newton, and other giantsof the 18th and 19th centuries, such as Bernoulli, Euler, Lagrange, and Laplace, pro-vided the building blocks in solid mechanics, fluid mechanics, and mathematics. Theneeded applications to aeronautics were made mostly after 1900 by workers in manycountries, of whom special reference should be made to the Wright brothers, G. H.Bryan, F. W. Lanchester, J. C. Hunsaker, H. B. G1auert, B. M. Jones, and S. B. Gates.These pioneers introduced and extended the basis for analysis and experiment thatunderlies all modem practice," This body of knowledge is well documented in severaltexts of that period, for example, Bairstow (1939). Concurrently, principally in theUnited States of America and Britain, a large body of aerodynamic data was accumu-lated, serving as a basis for practical design.

Newton's laws of motion provide the connection between environmental forcesand resulting motion for all but relativistic and quantum-dynamical processes, includ-ing all of "ordinary" and much of celestial mechanics. What then distinguishes flightdynamics from other branches of applied mechanics? Primarily it is the special na-ture of the force fields with which we have to be concerned, the absence of the kine-matical constraints central to machines and mechanisms, and the nature of the controlsystems used in flight. The external force fields may be identified as follows:

"Strong" Fields:1. Gravity2. Aerodynamic3. Buoyancy

"Weak" Fields:4. Magnetic5. Solar radiation

We should observe that two of these fields, aerodynamic and solar radiation, pro-duce important heat transfer to the vehicle in addition to momentum transfer (force).Sometimes we cannot separate the thermal and mechanical problems (Etkin andHughes, 1967). Of these fields only the strong ones are of interest for atmosphericand oceanic flight, the weak fields being important only in space. It should be re-marked that even in atmospheric flight the gravity force can not always be approxi-mated as a constant vector in an inertial frame. Rotations associated with Earth cur-vature, and the inverse square law, become important in certain cases of high-speedand high-altitude flight (Etkin, 1972).

The prediction, measurement and representation of aerodynamic forces are theprincipal distinguishing features of flight dynamics. The size of this task is illustrated

2An excellent account of the early history is given in the 1970 von Karman Lecture by Perkins(1970).

Page 14: Stability Control EtkinReid

4 Chapter 1. Introduction

SHAPE: sectionsc:=:::::::... <:::::::>

Parametersof wing aerodynamics

o 1.0SPEED:I I

SubsonicIncompressible Transonic

Supersonic

5.0I ;;,oM

Hypersonic

MOTION: Constant velocitylu, II, W,p, q, rIa const

Variablevelocitylu(t), v(t), w(t), p(t), q(t), ret»)

ATMOSPHERE:r-j -----,jr-------,jContinuum Slip Free-molecule

IUniform and

at rest

j INonuniform and Uniform and

at rest in motion(reentry) (gusts)

Spectrum of aerodynamic problems for wings.

jNonuniform and

in motion

Figure 1.2

by Fig. 1.2, which shows the enormous range of variables that need to be consideredin connection with wings alone. To be added, of course, are the complications ofpropulsion systems (propellers, jets, rockets), compound geometries (wing + body +tail), and variable geometry (wing sweep, camber).

As remarked above, Newton's laws state the connection between force and mo-tion. The commonest problem consists of finding the motion when the laws for theforces are given (all the numerical examples given in this book are of this kind).However, we must be aware of certain important variations:

.•

1. Inverse problems of first kind-the system and the motion are given and theforces have to be calculated.

2. Inverse problems of the second kind-the forces and the motion are given andthe system parameters have to be found.

3. Mixed problems-the unknowns are a mixture of variables from the force,system, and motion.

Examples of these inverse and mixed problems often tum up in research, whenone is trying to deduce aerodynamic forces from the observed motion of a vehicle inflight or of a model in a wind tunnel. Another example is the deduction of harmonicsof the Earth's gravity field from observed perturbations of satellite orbits. Theseproblems are closely related to the "plant identification" or "parameter identification"problem of system theory. [Inverse problems were treated in Chap. 11 of Etkin(1959)].

Page 15: Stability Control EtkinReid

1.2 The Tools of Flight Dynamicists 5

TYPES OF PROBLEMS

The main types of flight dynamics problem that occur in engineering practice are:

1. Calculation of "performance" quantities, such as speed, height, range, andfuel consumption.

2. Calculation of trajectories, such as launch, reentry, orbital and landing.3. Stability of motion.4. Response of vehicle to control actuation and to propulsive changes.5. Response to atmospheric turbulence, and how to control it.6. Aeroelastic oscillations (flutter).7. Assessment of human-pilot/machine combination (handling qualities).

It takes little imagination to appreciate that, in view of the many vehicle typesthat have to be dealt with, a number of subspecialties exist within the ranks of flightdynamicists, related to some extent to the above problem categories. In the context ofthe modern aerospace industry these problems are seldom simple or routine. On thecontrary they present great challenges in analysis, computation, and experiment.

1.2 The Toolsof Flight DynamicistsThe tools used by flight dynamicists to solve the design and operational problems ofvehicles are of three kinds:

1. Analytical2. Computational3. Experimental

The analytical tools are essentially the same as those used in other branches ofmechanics, that is the methods of applied mathematics. One important branch of ap-plied mathematics is what is now known as system theory, including stability, auto-matic control, stochastic processes and optimization. Stability of the uncontrolled ve-hicle is neither a necessary nor a sufficient condition for successful controlled flight.Good airplanes have had slightly unstable modes in some part of their flight regime,and on the other hand, a completely stable vehicle may have quite unacceptable han-dling qualities. It is dynamic performance criteria that really matter, so to expend agreat deal of analytical and computational effort on finding stability boundaries ofnonlinear and time-varying systems may not be really worthwhile. On the other hand,the computation of stability of small disturbances from a steady state, that is, the lin-ear eigenvalue problem that is normally part of the system study, is very useful in-deed, and may well provide enough information about stability from a practicalstandpoint.

On the computation side, the most important fact is that the availability of ma-chine computation has revolutionized practice in this subject over the past fewdecades. Problems of system performance, system design, and optimization that

Page 16: Stability Control EtkinReid

6 Chapter 1. Introduction

could not have been tackled at all in the past are now handled on a more or less rou-tine basis.

The experimental tools of the flight dynamicist are generally unique to this field.First, there are those that are used to find the aerodynamic inputs. Wind tunnels andshock tubes that cover most of the spectrum of atmospheric flight are now availablein the major aerodynamic laboratories of the world. In addition to fixed laboratoryequipment, there are aero ballistic ranges for dynamic investigations, as well asrocket-boosted and gun-launched free-flight model techniques. Hand in hand with thedevelopment of these general facilities has gone that of a myriad of sensors and in-struments, mainly electronic, for measuring forces, pressures, temperatures, accelera-tion, angular velocity, and so forth. The evolution of computational fluid dynamics(CFD) has sharply reduced the dependence of aerodynamicists on experiment. Manyresults that were formerly obtained in wind tunnel tests are now routinely providedby CFD analyses. The CFD codes themselves, of course, must be verified by compar-ison with experiment.

Second, we must mention the flight simulator as an experimental tool used di-rectly by the flight dynamicist. In it he studies mainly the matching of the pilot to themachine. This is an essential step for radically new flight situations. The ability of thepilot to control the vehicle must be assured long before the prototype stage. This can-not yet be done without test, although limited progress in this direction is being madethrough studies of mathematical models of human pilots. Special simulators, built formost new major aircraft types, provide both efficient means for pilot training, and aresearch tool for studying handling qualities of vehicles and dynamics of human pi-lots. The development of high-fidelity simulators has made it possible to greatly re-duce the time and cost of training pilots to fly new types of airplanes.

,

••

,1.3 Stability, Control, and Equilibrium

It is appropriate here to define what is meant by the terms stability and control. To doso requires that we begin with the concept of equilibrium.

A body is in equilibrium when it is at rest or in uniform motion (i.e., has constantlinear and angular momenta). The most familiar examples of equilibrium are thestatic ones; that is, bodies at rest. The equilibrium of an airplane in flight, however, isof the second kind; that is, uniform motion. Because the aerodynamic forces are de-pendent on the angular orientation of the airplane relative to its flight path, and be-cause the resultant of them must exactly balance its weight, the equilibrium state iswithout rotation; that is, it is a motion of rectilinear translation.

Stability, or the lack of it, is a property of an equilibrium state." The equilibriumis stable if, when the body is slightly disturbed in any of its degrees of freedom, it re-turns ultimately to its initial state. This is illustrated in Fig. 1.3a. The remainingsketches of Fig. 1.3 show neutral and unstable equilibrium. That in Fig. 1.3d is amore complex kind than that in Fig. l.3b in that the ball is stable with respect to dis-placement in the y direction, but unstable with respect to x displacements. This has itscounterpart in the airplane, which may be stable with respect to one degree of free-dom and unstable with respect to another. Two kinds of instability are of interest in

31tis also possible to speak of the stability of a transient with prescribed initial condition.

Page 17: Stability Control EtkinReid

1.3 Stability, Control, and Equilibrium 7

(sl (bl (el

(dl

Figure 1.3 (a) Ball in a bowl-stable equilibrium. (b) Ball on a hill-unstable equilibrium. (c)Ball on a plane-neutral equilibrium. (d) Ball on a saddle surface-unstable equilibrium.

airplane dynamics. In the first, called static instability, the body departs continuouslyfrom its equilibrium condition. That is how the ball in Fig. 1.3b would behave if dis-turbed. The second, called dynamic instability, is a more complicated phenomenon inwhich the body oscillates about its equilibrium condition with ever-increasing ampli-tude.

When applying the concept of stability to airplanes, there are two classes thatmust be considered-inherent stability and synthetic stability. The discussion of theprevious paragraph implicitly dealt with inherent stability, which is a property ofthe basic airframe with either fixed or free controls, that is, control-fixed stability orcontrol-free stability. On the other hand, synthetic stability is that provided by an au-tomatic flight control system (APCS) and vanishes if the control system fails. Suchautomatic control systems are capable of stabilizing an inherently unstable airplane,or simply improving its stability with what is known as stability augmentation sys-tems (SAS). The question of how much to rely on such systems to make an airplaneflyable entails a trade-off among weight, cost, reliability, and safety. If the SASworks most of the time, and if the airplane can be controlled and landed after it hasfailed, albeit with diminished handling qualities, then poor inherent stability may beacceptable. Current aviation technology shows an increasing acceptance of SAS in allclasses of airplanes.

If the airplane is controlled by a human pilot, some mild inherent instability canbe tolerated, if it is something the pilot can control, such as a slow divergence. (Un-stable bicycles have long been ridden by humans!). On the other hand, there is no

Page 18: Stability Control EtkinReid

8 Chapter 1. Introduction

margin for error when the airplane is under the control of an autopilot, for then theclosed loop system must be stable in its response to atmospheric disturbances and tocommands that come from a navigation system.

In addition to the role controls play in stabilizing an airplane, there are two oth-ers that are important. The first is to fix or to change the equilibrium condition (speedor angle of climb). An adequate control must be powerful enough to produce thewhole range of equilibrium states of which the airplane is capable from a perfor-mance standpoint. The dynamics of the transition from one equilibrium state to an-other are of interest and are closely related to stability. The second function of thecontrol is to produce nonequilibrium, or accelerated motions; that is, maneuvers.These may be steady states in which the forces and accelerations are constant whenviewed from a reference frame fixed to the airplane (for example, a steady turn), orthey may be transient states. Investigations of the transition from equilibrium to anonequilibrium steady state, or from one maneuvering steady state to another, formpart of the subject matter of airplane control. Very large aerodynamic forces may acton the airplane when it maneuvers-a knowledge of these forces is required for theproper design of the structure.

RESPONSE TO ATMOSPHERIC TURBULENCEA topic that belongs in dynamics of flight and that is closely related to stability is theresponse of the airplane to wind gradients and atmospheric turbulence (Etkin, 1981).This response is important from several points of view. It has a strong bearing on theadequacy of the structure, on the safety of landing and take-off, on the acceptabilityof the airplane as a passenger transport, and on its accuracy as a gun or bombing plat-form.

1.4 TheHuman PilotAlthough the analysis and understanding of the dynamics of the airplane as an iso-lated unit is extremely important, one must be careful not to forget that for manyflight situations it is the response of the total system, made up of the human pilot andthe aircraft, that must be considered. It is for this reason that the designers of aircraftshould apply the findings of studies into the human factors involved in order to en-sure that the completed system is well suited to the pilots who must fly it.

Some of the areas of consideration include:

1. Cockpit environment; the occupants of the vehicle must be provided withoxygen, warmth, light, and so forth, to sustain them comfortably.

2. Instrument displays; instruments must be designed and positioned to providea useful and unambiguous flow of information to the pilot.

3. Controls and switches; the control forces and control system dynamics mustbe acceptable to the pilot, and switches must be so positioned and designed asto prevent accidental operation. Tables 1.1 to 1.3 present some pilot data con-cerning control forces.

4. Pilot workload; the workload of the pilot can often be reduced through properplanning and the introduction of automatic equipment.

Page 19: Stability Control EtkinReid

1.4 The Human Pilot 9

Table 1.1Estimates of the Maximum Rudder Forces that Can Be Exerted for Various Positions of theRudder Pedal (BuAer, 1954)

Rudder Pedal Position Distance from Back of Seat Pedal Force

(in) (em) (th) (N)

Back 31.00 78.74 246 1,094Neutral 34.75 88.27 424 1,886Forward 38.50 97.79 334 1,486

Table 1.2Hand-Operated Control Forces (From Flight Safety Foundation Human Engineering Bulletin56-5H) (see figure in Table 1.3)

Direction of Movement 180° 150° 120° 90° 60°

Rt. hand 52 56 42 37 24(231) (249) (187) (165) (107)

PullLft. hand 50 42 34 32 26

(222) (187) (151) (142) (116)

Rt. hand 50 42 36 36 34(222) (187) (160) (160) (151)

PushLft. hand 42 30 26 22 22

(187) (133) (116) (98) (98)

Rt. hand 14 18 24 20 20 Values given(62) (80) (107) (89) (89) represent

UpLft. hand 9 15

maximum17 17 15 exertable

(40) (67) (76) (76) (67) force inRt. hand 17 20 26 26 20 pounds

(76) (89) (116) (116) (89) (Newtons)Down by the 5

Lft. hand 13 18 21 21 18 percentile(58) (80) (93) (93) (80) man.

Rt. hand 14 15 15 16 17(62) (67) (67) (71) (76)

OutboardLft. hand 8 8 10 10 12

(36) (36) (44) (44) (53)

Rt. hand 20 20 22 18 20(89) (89) (98) (80) (89)

InboardLft. hand 13 15 20 16 17

(58) (67) (89) (71) (76)

Note: The above results are those obtained from unrestricted movement of the subject. Any force required toovercome garment restriction would reduce the effective forces by the same amount.

Page 20: Stability Control EtkinReid

10 Chapter 1. Introduction

DmECTION OF MOVEMENT

Vert. ref. line

lSO°

••

goo

Maximum Stick Average Rate of Stick Time for FullPull-up Load Motion Deflection ~

(lb) (N) (in/s) (cm/s) (s) ,,

1 35 156 51.85 131.70 0.1622 74 329 15.58 39.57 0.4753 77 343 11.00 27.94 0.6004 97 431 10.27 26.09 0.750

Outboard

tInboard

tOutboard

Table 1.3Rates of Stick Movement in Flight Test Pull-ups Under Various Loads (BuAer, 1954)

Page 21: Stability Control EtkinReid

1.5 Handling Qualities Requirements 11

The care exercised in considering the human element in the closed-loop systemmade up of pilot and aircraft can determine the success or failure of a given aircraftdesign to complete its mission in a safe and efficient manner.

Many critical tasks performed by pilots involve them in activities that resemblethose of a servo control system. For example, the execution of a landing approachthrough turbulent air requires the pilot to monitor the aircraft's altitude, position, atti-tude, and airspeed and to maintain these variables near their desired values throughthe actuation of the control system. It has been found in this type of control situationthat the pilot can be modeled by a linear control system based either on classical con-trol theory or optimal control theory (Etkin, 1972; Kleinman et al., 1970; McRuerand Krendel, 1973).

1.5 Handling Qualities RequirementsAs a result of the inability to carry out completely rational design of the pilot-machine combination, it is customary for the government agencies responsible forthe procurement of military airplanes, or for licensing civil airplanes, to specify com-pliance with certain "handling (or flying) qualities requirements" (e.g., ICAO, 1991;USAF, 1980; USAF, 1990). Handling qualities refers to those qualities or character-istics of an aircraft that govern the ease and precision with which a pilot is able toperform the tasks required in support of an aircraft role (Cooper and Harper, 1969).

These requirements have been developed from extensive and continuing flightresearch. In the final analysis they are based on the opinions of research test pilots,substantiated by careful instrumentation. They vary from country to country and fromagency to agency, and, of course, are different for different types of aircraft. They aresubject to continuous study and modification in order to keep them abreast of the lat-est research and design information. Because of these circumstances, it is not feasibleto present a detailed description of such requirements here. The following is intendedto show the nature, not the detail, of typical handling qualities requirements." Most ofthe specific requirements can be classified under one of the following headings.

CONTROL POWER

The term control power is used to describe the efficacy of a control in producing arange of steady equilibrium or maneuvering states. For example, an elevator control,which by taking positions between full up and full down can hold the airplane inequilibrium at all speeds in its speed range, for all configurations" and CG positions,is a powerful control. On the other hand, a rudder that is not capable at full deflectionof maintaining equilibrium of yawing moments in a condition of one engine out andnegligible sideslip is not powerful enough. The handling qualities requirements nor-mally specify the specific speed ranges that must be achievable with full elevator de-

4For a more complete discussion, see AGARD (1959); Stevens and Lewis (1992).sThis word describes the position of movable elements of the airplane-for example, landing con-

figuration means that landing flaps and undercarriage are down, climb configuration means that landinggear is up, and flaps are at take-off position, and so forth.

Page 22: Stability Control EtkinReid

12 Chapter 1. Introduction

flection in the various important configurations and the asymmetric power conditionthat the rudder must balance. They may also contain references to the elevator anglesrequired to achieve positive load factors, as in steady turns and pull-up maneuvers(see "elevator angle per g," Sec. 3.1).

CONTROL FORCES

The requirements invariably specify limits on the control forces that must be exertedby the pilot in order to effect specific changes from a given trimmed condition, or tomaintain the trim speed following a sudden change in configuration or throttle set-ting. They frequently also include requirements on the control forces in pull-up ma-neuvers (see "control force per g," Sec. 3.1). In the case of light aircraft, the controlforces can result directly from mechanical linkages between the aerodynamic controlsurfaces and the pilot's flight controls. In this case the hinge moments of Sec. 2.5playa direct role in generating these forces. In heavy aircraft, systems such as partialor total hydraulic boost are used to counteract the aerodynamic hinge moments and arelated or independent subsystem is used to create the control forces on the pilot'sflight controls.

STATIC STABILITY

The requirement for static longitudinal stability (see Chap. 2) is usually stated interms of the neutral point. The neutral point, defined more precisely in Sec. 2.3, is aspecial location of the center of gravity (CG) of the airplane. In a limited sense it isthe boundary between stable and unstable CG positions. It is usually required that therelevant neutral point (stick free or stick fixed) shall lie some distance (e.g., 5% ofthe mean aerodynamic chord) behind the most aft position of the CG. This ensuresthat the airplane will tend to fly at a constant speed and angle of attack as long as thecontrols are not moved.

The requirement on static lateral stability is usually mild. It is simply that thespiral mode (see Chap. 6) if divergent shall have a time to double greater than somestated minimum (e.g., 4s).

DYNAMIC STABILITY

The requirement on dynamic stability is typically expressed in terms of the dampingand frequency of a natural mode. Thus the USAF (1980) requires the damping andfrequency of the lateral oscillation for various flight phases and stability levels toconform to the values in Table 1.4.

STALLING AND SPINNING

Finally, most requirements specify that the airplane's behavior following a stall or ina spin shall not include any dangerous characteristics, and that the controls must re-tain enough effectiveness to ensure a safe recovery to normal flight.

",

Page 23: Stability Control EtkinReid

1.5 Handling Qualities Requirements 13

Table 1.41

Minimum Dutch Roll Frequency and Damping

Flight Phase Mintdwv * Min wnd'

Level Category Class Min td* rad/s rad/s

A I,IV 0.19 0.35 1.0n.m 0.19 0.35 0.4

1 B All 0.08 0.15 0.4

C I, II-C, IV 0.08 0.15 1.0

II-L,rn 0.08 0.10 0.4

2 All All 0.02 0.05 0.4

3 All All 0 - 0.4

'Level, Phase and Class are defined in USAF, 1980.*Note: The damping coefficient (, and the undamped natural frequency w•• are defined in Chap. 6.

RATING OF HANDLING QUALITIES

To be able to assess aircraft handling qualities one must have a measuring techniquewith which any given vehicle's characteristics can be rated. In the early days of avia-tion, this was done by soliciting the comments of pilots after they had flown the air-craft. However, it was soon found that a communications problem existed with pilotsusing different adjectives to describe the same flight characteristics. These ambigui-ties have been alleviated considerably by the introduction of a uniform set of descrip-tive phrases by workers in the field. The most widely accepted set is referred to as the"Cooper-Harper Scale," where a numerical rating scale is utilized in conjunction witha set of descriptive phrases. This scale is presented in Fig. 1.4. To apply this ratingtechnique it is necessary to describe accurately the conditions under which the resultswere obtained. In addition it should be realized that the numerical pilot rating (1-10)is merely a shorthand notation for the descriptive phrases and as such no mathemati-cal operations can be carried out on them in a rigorous sense. For example, a vehicleconfiguration rated as 6 should not be thought to be "twice as bad" as one rated at 3.The comments from evaluation pilots are extremely useful and this information willprovide the detailed reasons for the choice of a rating.

Other techniques have been applied to the rating of handling qualities. For exam-ple, attempts have been made to use the overall system performance as a rating pa-rameter. However, due to the pilot's adaptive capability, quite often he can cause theoverall system response of a bad vehicle to approach that of a good vehicle, leadingto the same performance but vastly differing pilot ratings. Consequently system per-formance has not proved to be a good rating parameter. A more promising approachinvolves the measurement of the pilot's physiological and psychological state. Suchmethods lead to objective assessments of how the system is influencing the humancontroller. The measurement of human pilot describing functions is part of this tech-nique (Kleinman et al., 1970; McRuer and Krendel, 1973; Reid, 1969).

Research into aircraft handling qualities is aimed in part at ascertaining whichvehicle parameters influence pilot acceptance. It is obvious that the number of possi-

Page 24: Stability Control EtkinReid

ADEQUACY FOR SELECTED TASK OR AIRCRAFT CHARACTERISTICS • DEMANDS ON THE PILOT PILOT JREQUIRED OPERATION* IN SELECTED TASK OR REQUIRED OPERATION* RATING

Excellent • Pilot compensation not a factor for 1Highly desirable desired performance

Good • Pilot compensation not a factor for2

Negligible deficiencies desired performance

Fair - Some mildly • Minimal pilot compensation required for 3unpleasant deficiencies desired performance

Yes Minor but annoying • Desired performance requires moderate4deficiencies pilot compensation

Is itNo Deficiencies

satisfactory without Moderately objectionable Adequate performance requires considerablewarrant r---- • 5improvement? improvement deficiencies pilot compensation

Very objectionable but • Adequate performance requires extensive6

tolerable deficiencies pilot compensationYes

Adequate performance not attainable withMajor deficiencies • maximum tolerable pilot compension. 7

Is adequate Controllability not in question.performance No Deficiencies

attainable with a require r--- Considerable pilot compensation is requiredMajor deficiencies • 8tolerable pilot improvement for control

workload

Major deficiencies • Intense pilot compensation is required to 9retain control

Yes

Is it No ( Improvement Major deficienciesControl will be lost during some portion of

10controllable? l mandatory • required operation.

( Pilot * Definition of required operation involves designation of flightdecisions phase and/or subphase with accompanying conditions.

Figure 1.4 Handling qualities rating scale; CooperlHarper scale (Cooper and Harper, 1969).

Page 25: Stability Control EtkinReid

1.6 Axes andNotation 15

6.0 Initial response fast,oversensitive, light

slick fon:es

OL.....--------I-----I~-__:_'_:-~~~0.1 0.5. 1.0 2.0 3.0 4.0

Damping ratio, rFigure 1.5 Longitudinal short-period oscillation-pilot opinion contours (O'Hara, 1967).

IniUal response fast,tendency tooscillation andto 0VIISh00tloads

.......--.....,\

S1uglsh, largestlck IIIlIIIonand fon:es

Acc:aptable

ble combinations of parameters is staggering, and consequently attempts are made tostudy one particular aspect of the vehicle while maintaining all others in a "satisfac-tory" configuration. Thus the task is formulated in a fashion that is amenable tostudy. The risk involved in this technique is that important interaction effects can beoverlooked. For example, it is found that the degree of difficulty a pilot finds in con-trolling an aircraft's lateral-directional mode influences his rating of the longitudinaldynamics. Such facts must be taken into account when interpreting test results. An-other possible bias exists in handling qualities results obtained in the past becausemost of the work has been done in conjunction with fighter aircraft. The findingsfrom such research can often be presented as "isorating" curves such as those shownin Fig. 1.5.

Unacceptable

V8IY slow response,large control motionto maneuver,diffICUlt to trim

In this book the Earth is regarded as flat and stationary in inertial space. Any coordi-nate system, or frame of reference, attached to the Earth is therefore an inertial sys-tem, one in which Newton's laws are valid. Clearly we shall need such a referenceframe when we come to formulate the equations of motion of a flight vehicle. We de-note that frame by FE(OE,xE,yE,ZE)' Its origin is arbitrarily located to suit the circum-stances of the problem, the axis OEZE points vertically downward, and the axis 0JiXE'which is horizontal, is chosen to point in any convenient direction, for example,North, or along a runway, or in some reference flight direction. It is additionally as-sumed that gravity is uniform, and hence that the mass center and center of gravity(CG) are the same point. The location of the CG is given by its Cartesian coordinatesrelative to FE' Its velocity relative to FE is denoted VE and is frequently termed thegroundspeed.

1.6 Axes and Notation

Page 26: Stability Control EtkinReid

16 Chapter 1. Introduction

Z,W

II

Figure 1.6 Notation for body axes. L = rolling moment, M = pitching moment, N = yawingmoment, p = rate of roll, q = rate of pitch, r = rate of yaw. [X, Y, Z] = components of resultantaerodynamic force. [u, v, w] = components of velocity of C relative to atmosphere.

Aerodynamic forces, on the other hand, depend not on the velocity relative to FE'but rather on the velocity relative to the surrounding air mass (the airspeed), whichwill differ from the groundspeed whenever there is a wind. If we denote the wind ve-locity vector relative to FE by W, and that of the CG relative to the air by V thenclearly

VE=V+W (1.6,1)

The components of W in frame FE' that is, relative to Earth, are given by

WE = [Wx w, wzt (1.6,2)

V represents the magnitude of the airspeed (thus retaining the usual aerodynamicsmeaning of this symbol). For the most part we will have W = 0, making the airspeedthe same as the inertial velocity.

A second frame of reference will be needed in the development of the equationsof motion. This frame is fixed to the airplane and moves with it, having its origin C atthe CG, (see Fig. 1.6). It is denoted FB and is commonly called body axes. Cxz is theplane of symmetry of the vehicle. The components of the aerodynamic forces andmoments that act on the airplane, and of its linear and angular velocities relative to

y

1/

~or--~"'-__----

Projection of V on x% plane

lJW ~Figure 1.7 (a) Definition of ax. (b) View in plane ofy and V, definition of f3.

Page 27: Stability Control EtkinReid

1.6 Axes and Notation 17

the air are denoted by the symbols given in the figure. In the notation of AppendixA.I, this means, for example, that

VB = [uvwY (1.6,3)

The vector V does not in general lie in any of the coordinate planes. Its orienta-tion is defined by the two angles shown in Fig. 1.7:

Angle of attack,

(1.6,4)

Angle of sideslip,v

f3 = sin-1 -V

With these definitions, the sideslip angle f3 is not dependent on the direction of ex inthe plane of symmetry.

The symbols used throughout the text correspond generally to current usage andare mainly used in a consistent manner.

Page 28: Stability Control EtkinReid

CHAPTER 2

Static Stability and Control-Part]

2.1 General RemarksA general treatment of the stability and control of airplanes requires a study of thedynamics of flight, and this approach is taken in later chapters. Much useful informa-tion can be obtained, however, from a more limited view, in which we consider notthe motion of the airplane, but only its equilibrium states. This is the approach inwhat is commonly known as static stability and control analysis.

The unsteady motions of an airplane can frequently be separated for convenienceinto two parts. One of these consists of the longitudinal or symmetric motions; that is,those in which the wings remain level, and in which the center of gravity moves in avertical plane. The other consists of the lateral or asymmetric motions; that is,rolling, yawing, and sideslipping, while the angle of attack, the speed, and the angleof elevation of the x axis remains constant.

This separation canbe made for both dynamic and static analyses. However, theresults of greatest importance for static stability are those associated with the longitu-dinal analysis. Thus the principal subject matter of this and the following chapter isstatic longitudinal stability and control. A brief discussion of the static aspects of di-rectional and rolling motions is contained in Sees, 3.9 and 3.11.

We shall be concerned with two aspects of the equilibrium state. Under the head-ing stability we shall consider the pitching moment that acts on the airplane when itsangle of attack is changed from the equilibrium value, as by a vertical gust. We focusour attention on whether or not this moment acts in such a sense as to restore the air-plane to its original angle of attack. Under the heading control we discuss the use of alongitudinal control (elevator) to change the equilibrium value of the angle of attack.

The restriction to angle of attack disturbances when dealing with stability mustbe noted, since the applicability of the results is thereby limited. When the aerody-namic characteristics of an airplane change with speed, owing to compressibility ef-fects, structural distortion, or the influence of the propulsive system, then the airplanemay be unstable with respect to disturbances in speed. Such instability is not pre-dicted by a consideration of angle of attack disturbances only. (See Fig. 1.3d, andidentify speed with x, angle of attack with y.) A more general point of view than thatadopted in this chapter is required to assess that aspect of airplane stability. Such aviewpoint is taken in Chap. 6. To distinguish between true general static stability andthe more limited version represented by em vs. a, we use the term pitch stiffness forthe latter.

18

Page 29: Stability Control EtkinReid

2.1 General Remarks 19

Although the major portion of this and the following chapter treats a rigid air-plane, an introduction to the effects of airframe distortion is contained in Sec. 3.5.

THE BASIC LONGITUDINAL FORCES

The basic flight condition for most vehicles is symmetric steady flight. In this condi-tion the velocity and force vectors are as illustrated in Fig. 2.1. All the nonzero forcesand motion variables are members of the set defined as "longitudinal." The two mainaerodynamic parameters of this condition are V and a.

Nothing can be said in general about the way the thrust vector varies with V anda, since it is so dependent on the type of propulsion unit-rockets, jet, propeller, orturboprop. 1Woparticular idealizations are of interest, however,

1. T independent of V, that is, constant thrust; an approximation for rockets andpure jets.

2. TV independent of V, that is, constant power; an approximation for reciprocat-ing engines with constant-speed propellers.

The variation of steady-state lift and drag with a for subsonic and supersonicMach numbers (M < about 5) are characteristically as shown in Fig. 2.2 for therange of attached flow over the surfaces of the vehicle (McCormick, 1994; Miele,1962; Schlichting and Truckenbrodt, 1979). Over a useful range of a (below the stall)the coefficients are given accurately enough by

CL = c;« (2.1,1)CD = CDmin+ KCL

2 (2.1,2)

The three constants CLa, CDmin,K are principally functions of the configuration shape,thrust coefficient, and Mach number.

Significant departure from the above idealizations may, of course, be anticipatedin some cases. The minimum of CD may occur at a value of a> 0, and the curvatureof the CL vs. a relation may be an important consideration for flight at high CL'When the vehicle is a "slender body," for example, a slender delta, or a slim wingless

L

WFigure 2.1 Steady symmetric flight.

Page 30: Stability Control EtkinReid

20 Chapter 2. Static Stability and Control-Part 1

--+-------a --;--------,a

Figure 2.2 Lift and drag for subsonic and supersonic speeds.

body, the CL curve may have a characteristic upward curvature even at small a (Flaxand Lawrence, 1951). The upward curvature of CL at small a is inherently present athypersonic Mach numbers (Truitt, 1959). For the nonlinear cases, a suitable formula-tion for CL is (USAF, 1978)

CL = (~CNa sin 2a + CNaa sin alsin al) cos a (2.1,3)

where CNa and CNaa are coefficients (independent of a) that depend on the Machnumber and configuration. [Actually CN here is the coefficient of the aerodynamicforce component normal to the wing chord, and CNa is the value of CLa at a = 0, ascan easily be seen by linearizing (2.1,3) with respect to o.] Equation (2.1,2) for thedrag coefficient can serve quite well for flight dynamics applications up to hyper-sonic speeds (M > 5) at which theory indicates that the exponent of CL decreasesfrom 2 to i. Miele (1962) presents in Chap. 6 a very useful and instructive set oftypical lift and drag data for a wide range of vehicle types, from subsonic to hyper-sonic. •

Balance, or EquilibriumAn airplane can continue in steady unaccelerated flight only when the resultant

external force and moment about the CG both vanish. In particular, this requires thatthe pitching moment be zero. This is the condition of longitudinal balance. If thepitching moment were not zero, the airplane would experience a rotational accelera-tion component in the direction of the unbalanced moment. Figure 2.3 shows a typi-cal graph of the pitching-moment coefficient about the CG1 versus the angle of attackfor an airplane with a fixed elevator (curve a). The angle of attack is measured fromthe zero-lift line of the airplane. The graph is a straight line except near the stall.Since zero Cm is required for balance, the airplane can fly only at the angle of attackmarked A, for the given elevator angle.

I

IUnless otherwise specified, em always refers to moment about the CG.

Page 31: Stability Control EtkinReid

2.1 General Remarks 21

em

Balanced and positive stiffnessNose

up emo.."..__ ...--- b

_-B

Nosedown

---~--Balanced but negative stiffness

"Figure 2.3 Pitching moment of an airplane about the eG.

Pitch StiffnessSuppose that the airplane of curve a on Fig. 2.3 is disturbed from its equilibrium

attitude, the angle of attack being increased to that at B while its speed remains unal-tered. It is now subject to a negative, or nose-down, moment, whose magnitude corre-sponds to Be. This moment tends to reduce the angle of attack to its equilibriumvalue, and hence is a restoring moment. In this case, the airplane has positive pitchstiffness, obviously a desirable characteristic.

On the other hand, if Cm were given by the curve b, the moment acting when dis-turbed would be positive, or nose-up, and would tend to rotate the airplane still far-ther from its equilibrium attitude. We see that the pitch stiffness is determined by thesign and magnitude of the slope aC,jaa. If the pitch stiffness is to be positive at theequilibrium <x,Cmmust be zero, and dCrr/ih must be negative. It will be appreciatedfrom Fig. 2.3 that an alternative statement is "Cmo must be positive, and dCm/ih neg-ative if the airplane is to meet this (limited) condition for stable equilibrium." Thevarious possibilities corresponding to the possible signs of Cmo and aCmlaa areshown in Figs. 2.3 and 2.4.

1ii-

Figure 2.4 Other possibilities.

Page 32: Stability Control EtkinReid

22 Chapter 2. Static Stability and Control-Part 1

E--~ E- ~----~- ~---~Positive camber Zero camber Negative camberCillO negative CillO = 0 CillO positive

Figure 2.5 C"",of airfoil sections.

Possible ConfigurationsThe possible solutions for a suitable configuration are readily discussed in terms

of the requirements on Cmo and oC,joa. We state here without proof (this is given inSec. 2.3) that oC,joa can be made negative for virtually any combination of liftingsurfaces and bodies by placing the center of gravity far enough forward. Thus it is notthe stiffness requirement, taken by itself, that restricts the possible configurations, butrather the requirement that the airplane must be simultaneously balanced and havepositive pitch stiffness. Since a proper choice of the CG location can ensure a nega-tive oC,joa, then any configuration with a positive Cmo can satisfy the (limited) con-ditions for balanced and stable flight.

Figure 2.5 shows the Cmo of conventional airfoil sections. If an airplane were toconsist of a straight wing alone (flying wing), then the wing camber would determinethe airplane characteristics as follows:

Negative camber-flight possible at a > 0; i.e., CL > 0 (Fig. 2.3a).Zero camber-flight possible only at a = 0, or CL = O.Positive camber-flight not possible at any positive a or Cv

For straight-winged tailless airplanes, only the negative camber satisfies the con-ditions for stable, balanced flight. Effectively the same result is attained if a flap, de-flected upward, is incorporated at the trailing edge of a symmetrical airfoil. A con-ventional low-speed airplane, with essentially straight wings and positive camber,could fly upside down without a tail, provided the CG were far enough forward(ahead of the wing mean aerodynamic center). Flying wing airplanes based on astraight wing with negative camber are not in general use for three main reasons:

+ Cambered wing at CL = 0 Tail with CL negative

CG

(sl

tTail with CL positive + Cambered wing at CL = 0

(bl

Figure 2.6 Wing-tail arrangements with positive C"",.(a) Conventional arrangement. (b) Tail-firstor canard arrangement.

Page 33: Stability Control EtkinReid

2.2 Synthesis of Lift and Pitching Moment 23

+ Lift

- Lift

Figure 2.7 Swept-back wing with twisted tips.

1. The dynamic characteristics tend to be unsatisfactory.2. The permissible CG range is too small.3. The drag and CLmax characteristics are not good.

The positively cambered straight wing can be used only in conjunction with anauxiliary device that provides the positive Cmo' The solution adopted by experi-menters as far back as Samuel Henson (1842) and John Stringfellow (1848) was toadd a tail behind the wing. The Wright brothers (1903) used a tail ahead of the wing(canard configuration). Either of these alternatives can supply a positive Cmo' as illus-trated in Fig. 2.6. When the wing is at zero lift, the auxiliary surface must provide anose-up moment. The conventional tail must therefore be at a negative angle of at-tack, and the canard tail at a positive angle.

An alternative to the wing-tail combination is the swept-back wing with twistedtips (Fig. 2.7). When the net lift is zero, the forward part of the wing has positive lift,and the rear part negative. The result is a positive couple, as desired.

A variant of the swept-back wing is the delta wing. The positive Cmo can beachieved with such planfonns by twisting the tips, by employing negative camber, orby incorporating an upturned tailing edge flap.

2.2 Synthesis of Lift and Pitching MomentThe total lift and pitching moment of an airplane are, in general, functions of angle ofattack, control-surface angle(s), Mach number, Reynolds number, thrust coefficient,and dynamic pressure' (The last-named quantity enters because of aeroelastic ef-fects. Changes in the dynamic pressure (~py2), when all the other parameters are con-stant, may induce enough distortion of the structure to alter Cm significantly.) An ac-curate determination of the lift and pitching moment is one of the major tasks in astatic stability analysis. Extensive use is made of wind-tunnel tests, supplemented byaerodynamic and aeroelastic analyses.

2When partial derivatives are taken in the following equations with respect to one of these variables,for example, iJC,jiJa, it is to be understood that all the others are held constant.

Page 34: Stability Control EtkinReid

24 Chapter 2. Static Stability and Control-Part 1

For purposes of estimation, the total lift and pitching moment may be synthe-sized from the contributions of the various parts of the airplane, that is, wing, body,nacelles, propulsive system, and tail, and their mutual interferences. Some data for .estimating the various aerodynamic parameters involved are contained in AppendixB, while the general formulation of the equations, in terms of these parameters, fol-lows here. In this chapter aeroelastic effects are not included. Hence the analysis ap-plies to a rigid airplane.

LIFT AND PITCHING MOMENT OF THE WING

The aerodynamic forces on any lifting surface can be represented as a lift and dragacting at the mean aerodynamic center, together with a pitching couple independentof the angle of attack (Fig. 2.8). The pitching moment of this force system about theCG is given by (Fig. 2.9?

Mw = Macw + (Lw cos CYw + Dw sin cxw)(h - hn)c+ (Lw sin CXw- Dw cos cxw)z (2.2,1)

We assume that the angle of attack is sufficiently small to justify the approximations

cos CXw= 1,

and the equation is made nondimensional by dividing through by ipv2Sc. It then be-comes

c.; = Cmacw + (CLw + Cv."CYw)(h - hllJ + (CLwcxw - CvJzlc (2.2,2)

Although it may occasionally be necessary to retain all the terms in (2.2,2), experi-ence has shown that the last term is frequently negligible, and that Cvwcxw may be ne-glected in comparison with CLw. With these simplifications, we obtain

Figure 2.8 Aerodynamic forces on the wing.

3The notation hn indicates that the mean aerodynamic center of the wing is also the neutral point ofthe wing. Neutral poi~t is defined in Sec. 2.3.

Page 35: Stability Control EtkinReid

2.2 Synthesis of Lift and Pitching Moment 25

Mean lWingzero aerodynamic '"""-----lift direction cho:.;.rd=-_'-::L-H~:;""' +-__ ---l II'"5 t~,

V \ (8

I+------c----~

I+----hc---+j

Figure 2.9 Moment about the CG in the plane of symmetry.

(2.2,3)

where aw = CL is the lift-curve-slope of the wing.Equation £2,3 will be used to represent the wing pitching moment in the discus-

sions that follow.

LIFT AND PITCHING MOMENT OF THE BODY AND NACELLES

The influences of the body and nacelles are complex. A body alone in an airstream issubjected to aerodynamic forces. These, like those on the wing, may be representedover moderate ranges of angle of attack by lift and drag forces at an aerodynamiccenter, and a pitching couple independent of a. Also as for a wing alone, the lift-a re-lation is approximately linear. When the wing and body are put together, however, asimple superposition of the aerodynamic forces that act upon them separately doesnot give a correct result. Strong interference effects are usually present, the flow fieldof the wing affecting the forces on the body, and vice versa.

These interference flow fields are illustrated for subsonic flow in Fig. 2.10. Part(a) shows the pattern of induced velocity along the body that is caused by the wingvortex system. This induced flow produces a positive moment that increases withwing lift or a. Hence a positive (destabilizing) contribution to Cma results. Part (b)shows an effect of the body on the wing. When the body axis is at angle a to thestream, there is a cross-flow component V sin a. The body distorts this flow locally,leading to cross-flow components that can be of order 2V sin a at the wing-body in-tersection. There is a resulting change in the wing lift distribution.

The result of adding a body and nacelles to a wing may usually be interpreted asa shift (forward) of the mean aerodynamic center, an increase in the lift-curve slope,and a negative increment in Cmac' The equation that corresponds to (2.2,3) for a wing-body-nacelle combination is then of the same form as (2.2,3), but with different val-ues of the parameters. The subscript wb is used to denote these values.

CmWb = Cmacwb + CLwb(h - hnwb)

= Cmacwb + awbCXwb(h - hnwb)

where awb is the lift-curve-slope of the wing-body-nacelle combination.

(2.2,4)

Page 36: Stability Control EtkinReid

26 Chapter 2. Stalk Stability and Control-Part 1

(a)

V sin a

t Vsin a(6)

Figure 2.10 Example of mutual interference flow fields of wing and body-subsonic flow. (a)Qualitative pattern of upwash and downwash induced along the body axis by the wing vorticity. (b)Qualitative pattern of upwash induced along wing by the cross-flow past the body.

LIFT AND PITCHING MOMENT OF THE TAIL

The forces on an isolated tail are represented just like those on an isolated wing.When the tail is mounted on an airplane, however, important interferences occur. Themost significant of these, and one that is usually predictable by aerodynamic theory,is a downward deflection of the flow at the tail caused by the wing. This is character-ized by the mean downwash angle E. Blanking of part of the tail by the body is a sec-ond effect, and a reduction of the relative wind when the tail lies in the wing wake isthe third.

Figure 2.11 depicts the forces acting on the tail showing the relative wind vectorof the airplane. V' is the average or effective relative wind at the tail. The tail lift anddrag forces are, respectively, perpendicular and parallel to V'. The reader should note

Wing meanaerodynamicchord

r...c-G--------;-----l-"u>-b---e)----'~

-.1----------'-'--

Tail meanaerodynamicchord

Figure 2.11 Forces acting on the tail.

Page 37: Stability Control EtkinReid

2.2 Synthesis of Lift and Pitching Moment 27

the tail angle it, which must be positive as shown for equilibrium. This is sometimesreferred to as longitudinal dihedral.

The contribution of the tail to the airplane lift, which by definition is perpendicu-lar to V, is

L, cos E - D, sin E

E is always a small angle, and we assume that DtE may be neglected compared withLt. The contribution of the tail to the airplane lift then becomes simply Lt. We intro-duce the symbol CLr to represent the lift coefficient of the tail, based on the airplanedynamic pressure iplf'! and the tail area St.

LtCL, = ipV2S

t(2.2,5)

The total lift of the airplane is

or in coefficient form

s,CL = ci; + S CL,

The reader should note that the lift coefficient of the tail is often based on the localdynamic pressure at the tail, which differs from ipy2 when the tail lies in the wingwake. This practice entails carrying the ratio Y'/Y in many subsequent equations. Thedefinition employed here amounts to incorporating V'/V into the tail lift-curve slopeat. This quantity is in any event different from that for the isolated tail, owing to theinterference effects previously noted. This circumstance is handled in various ways inthe literature. Sometimes a tail efficiency factor 'TJt is introduced, the isolated tail liftslope being multiplied by 'TJt' In other treatments, 'TJt is used to represent (V'/Yi. Inthe convention adopted here, at is the lift-curve slope of the tail, as measured in situon the airplane, and based on the dynamic pressure ipy2• This is the quantity that isdirectly obtained in a wind-tunnel test.

From Fig. 2.11 we find the pitching moment of the tail about the CG to be

Mt = -It[Lt cos (awb - E) + Dt sin (awb - E)]

- Zt[Dt cos (awb - E) - L, sin (awb - E)] + Mac, (2.2,7)

Experience has shown that in the majority of instances the dominant term in thisequation is the first one, and that all others are negligible by comparison. Only thiscase will be dealt with here. The reader is left to extend the analysis to cases in whichthis approximation is not valid. With the above approximation, and that of small an-gles,

(2.2,6)

M, = -ltLt = -ltC L,ipIf'!S,

Upon conversion to coefficient form, we obtain

Mt It Stc; = ipy2Sc = - i S CL,

The combination ItS,ISc is the ratio of two volumes characteristic of the airplane's

(2.2,8)

Page 38: Stability Control EtkinReid

28 Chapter 2. Static Stability and Control-Part 1

Wing-bodymean aerodynamic

center

~------_'i'-------~------1,

Tail meanaerodynamic

center

Figure 2.12 Wing-body and tail mean aerodynamic centers.

geometry. It is commonly called the "horizontal-tail volume ratio," or more simply,the "tail volume." It is denoted here by VH' Thus

(2.2,9)

Since the center of gravity is not a fixed point, but varies with the loading condi-tion and fuel consumption of the vehicle, VH in (2.2,9) is not a constant (although itdoes not vary much). It is a little more convenient to calculate the moment of the tailabout a fixed point, the mean aerodynamic center of the wing-body combination, andto use this moment in the subsequent algebraic manipulations. Figure 2.12 shows therelevant relationships, and we define

(2.2,10)

which leads to

(2.2,11)

The moment of the tail about the wing-body mean aerodynamic center is then (cf.(2.2,9)]

em, = -VeCLt

and its moment about the CG is, from substitution of (2.2,11) into (2.2,9)

- Stc; = -VHCL, + CLtS (h - hnWb)

(2.2,12)

(2.2,13)

PITCHING MOMENT OF A PROPULSIVE SYSTEM

The moment provided by a propulsive system is in two parts: (1) that coming fromthe forces acting on the unit itself, for example, the thrust and in-plane force actingon a propeller, and (2) that coming from the interaction of the propulsive slipstreamwith the other parts of the airplane. These are discussed in more detail in Sec. 3.4. Weassume that the interference part is included in the moments already given for thewing, body, and tail, and denote by Cmp the remaining moment from the propulsionunits.

Page 39: Stability Control EtkinReid

2.3 Total Pitching Moment and Neutral Point 29

2.3 TotalPitching Moment and Neutral PointOn summing the first of (2.2,4) and (2.2,13) making use of (2.2,6) and adding thecontribution Cmp for the propulsive system, we obtain the total pitching momentabout the CG

C; = CmaCWb+ CL(h - hnwb) - fecL, + Cmp (2.3,1)

It is worthwhile repeating that no assumptions about thrust, compressibility, or aero-elastic effects have been made in respect of (2.3,1). The pitch stiffness (- Cm,) is nowobtained from (2.3,1). Recall that the mean aerodynamic centers of the wing-bodycombination and of the tail are fixed points, so that

ac _ ac acC = macwb+C (h-h )-V --.!::!..+----.!!!J!...rna aD! La nwb e aD! aD! (2.3,2)

If a true mean aerodynamic center in the classical sense exists, then aCmacw/aD! iszero and

(2.3,3)

Cma as given by (2.3,2) or (2.3,3) depends linearly on the CG position, h. Since CLa isusually large, the magnitude and sign of Cma depend strongly on h. This is the basisof the statement in Sec. 2.2 that Cma can always be made negative by a suitablechoice of h. The CG position hn for which Cma is zero is of particular significance,since this represents a boundary between positive and negative pitch stiffness. In thisbook we define h; as the neutral point, NP. It has the same significance for the vehi-cle as a whole as does the mean aerodynamic center for a wing alone, and indeed theterm vehicle aerodynamic center is an acceptable alternative to "neutral point."

The location of the NP is readily calculated from (2.3,2) by setting the left-handside to zero leading to

1 (aCmacwb - sc; aCmp)hn = hnWb- C aD! - Ve --a;; + a;;-

LaSubstitution of (2.3,4) back into (2.3,2) simplifies the latter to

(2.3,4)

(2.3,5)

which is valid whether Cmacwband Cmp vary with Cl or not. Equation (2.3,5) clearlyprovides an excellent way of finding hn from test results, that is from measurementsof Cma and CLa' The difference between the CG position and the NP is sometimescalled the static margin,

(2.3,6)

Since the criterion to be satisfied is Cma< 0, that is, positive pitch stiffness, thenwe see that we must have h < hm or K; > O. In other words the CG must be forwardof the NP. The farther forward the CG the greater is Km and in the sense of "staticstability" the more stable the vehicle.

The neutral point has sometimes been defined as the CG location at which thederivative dC,jdCL = O. When this definition is applied to the gliding flight of a rigid

Page 40: Stability Control EtkinReid

30 Chapter 2. Static Stability and Control-Part 1

airplane at low Mach number, the neutral point obtained is identical with that definedin this book. This is so because under these restricted conditions CL is a unique func-tion of a, and dC,jdCL == (dC,jda)ICoCJda). Then dC,jdCL and dC,jda are simul-taneously zero. In general, however, Cm and CL are both functions of several vari-ables, as pointed out at the beginning of Sec. 2.2. For fixed values of Se and h, andneglecting Reynolds number effects (these are usually very small), we may write

CL == f(a, M, C1"J~pV2), C; == gia, M, C1"J!pV2) (2.3,7)

where CT is the thrust coefficient, defined in Sec. 2.11.Mathematically speaking, the derivative dC,jdCL does not exist unless M, C1"J

and ~pyl are functions of CL.When that is the case, then

ac; dCm da sc; dM dCm dCT sc; a(ipV2)- == - - + - - + - - + (2.3,8)dCL da aCL dM aCL dCT aCL a(ipyl) dCL

Equation 2.3,8 has meaning only when a specific kind of flight is prescribed: e.g.,horizontal unaccelerated flight, or rectilinear climbing flight at full throttle. When acondition of this kind is imposed, then M, C1"Jand the dynamic pressure are definitefunctions of Co dC,jdCL exists, and a neutral point may be calculated. The neutralpoint so found is not an index of stability with respect to angle of attack disturbances,and the question arises as to what it does relate to. It can be shown that it relates tothe trim curves of the airplane. A plot of the elevator angle to trim versus speed willhave a zero slope when dC,jdCL is zero, and a negative slope when the CO lies aft ofthe neutral point so defined. As shown in Sec. 2.4, this reversal of slope indicates atendency toward instability with respect to speed, but only a dynamic analysis canshow whether or not the airplane is stable in this condition. There are cases when theapplication of the "trim-slope" criterion can be definitely misleading as to stability.One such is level unaccelerated flight, during which the throttle must be adjustedevery time the flight speed or CL is altered.

It can be seen from the foregoing remarks that the "trim-slope" criterion for theneutral point does not lead to any definite and clear-cut conclusions, either about thestability with respect to angle of attack disturbances, or about the general static sta-bility involving both speed and angle of attack disturbances. It is mainly for this rea-son that the neutral point has been defined herein on the basis of aC,jda.

EFFECT OF LINEAR LIFT AND MOMENT ON NEUTRAL POINT

When the forces and moments on the wing, body, tail, and propulsive system are lin-ear in a, as may be near enough the case in reality, some additional useful relationscan be obtained. We then have

and

CLwb = awbUwb

CLt = a.a,dCmC =C +-_P'a

mp mop da

(2.3,9)(2.3,10)

(2.3,11)

If CmWb is linear in CLwb, it can be shown (see exercise 2.3) that CmacWb does not varywith CLwb' i.e. that a true mean aerodynamic center exists. Figure 2.11 shows that thetail angle of attack is

Page 41: Stability Control EtkinReid

2.3 Total Pitching Moment and Neutral Point 31

(2.3,12)

(2.3,13)and hence CL, = arCOlwb - it - e)

The downwash e can usually be adequately approximated by

dee = Eo + dOl Olwb

The downwash eo at Olwb = 0 results from the induced velocity field of the body andfrom wing twist; the latter produces a vortex wake and downwash field even at zerototal lift. The constant derivative de/dOl occurs because the main contribution to thedownwash at the tail comes from the wing trailing vortex wake, the strength of whichis, in the linear case, proportional to CL-

The tail lift coefficient then is

(2.3,14)

CL, = a{ OlWb( 1 - ~:) - it - eo]and the total lift, from (2.2,6) and (2.3,9) is

(2.3,15)

or (b) (2.3,16)

or since Olwb and a differ by a constant

CL = aa (c)

where (2.3,17)

is the coefficient of lift when Olwb = 0;

a = dCL = awb[1 + !!!- St (1 _ :e)] (2.3,18)dOl awb S oa

is the lift-curve slope of the whole configuration; and Ol is the angle of attack of thezero-lift line of the whole configuration (see Fig. 2.13). Note that, since it is positive,

--+-+--------~ «wb

)0 «

Figure 2.13 Graph of total lift.

Page 42: Stability Control EtkinReid

32 Chapter 2. Static Stability and Control-Part 1

then (CL)o is negative. The difference between a and awb is found by equating(2.3,16b and c) to be

at S,a-rl/ =---(i+e)-.vb a S t 0 (2.3,19)

When the linear relations for CL> CLI and C""p are substituted into (2.3,1) the fol-lowing results can be obtained after some algebraic reduction:

CI1lQ = CmaCWb+ Cl1lQp+ at Va<eo+ it)- dC""CI1lQ = CI1lQ + (a - awb) ~

P P ua

at _ ( de) 1 dCmh =h +-V 1-- ----pn nwb a H da a da

Note that since C""o is the pitching moment at zero C¥wb,not at zero total lift, its valuedepends on h (via VH), whereas CI1lQ'being the moment at zero total lift, represents acouple and is hence independent of CG position. All the above relations apply to tail-less aircraft by putting VH = O. Another useful relation comes from integrating(2.3,5), i.e.

where

or

and

where

or

Cm = CI1lQ + CL(h - hn)Cm = CI1lQ + aaih - hn)

Cm", = a(h - hn)

(a)(2.3,20)

(b)

(a)

(2.3,21)(b)

(a)(2.3,22)

(b)

(c) ..(2.3,23)

(a)(b) (2.3,25)

(c)

Figure 2.14 shows the linear Cm vs. a relation, and Fig. 2.15 shows the resultant sys-tem of lift and moment that corresponds to (2.3,25), that is a force CL and a couple

em

Cm°tlllliii5S:::;===----------

Figure 2.14 Effect of CG location on em curve.

Page 43: Stability Control EtkinReid

2.4 Longitudinal Control 33

Figure 2.15 Total lift and moment acting on vehicle.

Cmo at the NP. Figure 2.15 is a very important result that the student should fix in hismind.

2.4 Longitudinal ControlIn this section we discuss the longitudinal control of the vehicle from a static point ofview. That is, we concern ourselves with how the equilibrium state of steady rectilin-ear flight is governed by the available controls. Basically there are two kinds ofchanges that can be made by the pilot or automatic control system-a change ofpropulsive thrust, or a change of configuration. Included in the latter are the opera-tion of aerodynamic controls--elevators, wing flaps, spoilers, and horizontal tail ro-tation. Since the equilibrium state is dominated by the requirement Cm = 0, the mostpowerful controls are those that have the greatest effect on Cm'

Figure 2.14 shows that another theoretically possible way of changing the trimcondition is to move the CG, which changes the value of a at which Cm = O.Movingit forward reduces the trim a or CL> and hence produces an increase in the trim speed.This method was actually used by Lilienthal, a pioneer of aviation, in gliding flightsduring 1891/1896, in which he shifted his body to move the CG. It has the inherentdisadvantage, apart from practical difficulties, of changing Cm", at the same time, re-ducing the pitch stiffness and hence stability, when the trim speed is reduced.

The longitudinal control now generally used is aerodynamic. A variable pitchingmoment is provided by moving the elevator, which may be all or part of the tail, or atrailing-edge flap in a tailless design. Deflection of the elevator through an angle Beproduces increments in both the Cm and CL of the airplane. The dCL caused by the el-evator of aircraft with tails is small enough to be neglected for many purposes. This isnot so for tailless aircraft, where the dCL due to elevators is usually significant. Weshall assume that the lift and moment increments for both kinds of airplane are linearin Be' which is a fair representation of the characteristics of typical controls at highReynolds number. Therefore,

dCL = CL8eBe (a)

CL = CL(a) + CL8eBe (b)(2.4,1)

dCm = Cm8eBe (c)

and Cm = Cm(a) + Cm8eBe (d)

where CL8e = dCddBe, Cm8e = dC,jdBe• and CL(a), Cm(a) are the "basic" lift and mo-ment when Be = O. The usual convention is to take down elevator as positive (Fig.2.16a). This leads to positive CL8e and negative Cm8e The deflection of the elevator

Page 44: Stability Control EtkinReid

34 Chapter 2. Static Stability and Control--Part 1

(6)

OJ-

.....I..L...-..__ •...•.. «o

(c)

Figure 2.16 Effect of elevator angle on Cm curve. (a) Elevator angle. (b) Cm - a curve. (c) CL -

a curve.

through a constant positive angle then shifts the Cm-a curve downward, withoutchange of slope (Fig. 2.16b). At the same time the zero-lift angle of the airplane isslightly changed (Fig. 2.16c).

In the case of linear lift and moment, we have

CL = CLaa + CLs/'eC; = Cmo + Cmaa + CmSe5e

(a)(2.4,2)

(b)

THE DERIVATIVES CLa.AND Cma.

Equation (2.2,6) gives the vehicle lift. Hence

dCL oc.; s, dCL,CLSe= as = ---as + S ase e e

(2.4,3)

Page 45: Stability Control EtkinReid

2.4 Longitudinal Control 3S

in which only the first term applies for tailless aircraft and the second for conven-tional tail elevators or all moving tails (when it is used instead of 0e)' We define theelevator lift effectiveness as

(2.4,4)

so that (2.4,3) becomes

(a) (2.4,5)

and the lift coefficient of the tail is

(b)

The total vehicle Cm is given for both tailed and tailless types by (2.3,1). For thelatter, of course, VH = O.Taking the derivative with respect to oe gives

oCC = macwb

mSe oOe (2.4,6)

We may usually neglect the last term, since there is unlikely to be any propulsive-ele-vator interaction that cannot be included in a.: Then (2.4,6) becomes

C oCmacwb + C (h h) -VmSe = 00 LSe - nwb - ae He

(2.4,7)

Summarizing for both types of vehicle, we have (retaining only the dominant terms)Tailed aircraft:

StCLSe = ae S

CmSe = -aeVH + CLSe(h - hnwb)

(a)(2.4,8)

(b)

Tailless aircraft:

(a)(2.4,9)

(b)

In the last case, the subscript wb is, of course, redundant and has been dropped. Theprimary parameters to be predicted or measured are a, for tailed aircraft, and oCJooe,oCmjooe for tailless.

ELEVATOR ANGLE TO TRIM

The trim condition is Cm = 0, whence from (2.4,1d)

8 = _ Cm(a)etrim Cml>e

(2.4,10)

Page 46: Stability Control EtkinReid

36 Chapter 2. Statie Stability and Control-Part 1

and the corresponding lift coefficient is

CLmm = CL(a) + CL6,,5etrim

(2.4,11)

When the linear lift and moment relations (2.4,2) apply the equations for trim are

(2.4,12)

These equations are solved for a and 8e to give

_ CmoCL6e + Cm6.CLrrimatrim - det

CmoCL", + Cm",CLmm8etrim = - det

(a)

(b) (2.4,13)

where

(c)

(d)

is the determinant of the square matrix in (2.4,12) and is normally negative. The val-ues of det for the two types of airplane are readily calculated from (2.4,8 and 9) to-gether with (2.3,5) to give

Tailed aircraft:

(a)

Tailless aircraft:

(b) (2.4,14)

and both are independent of h. From (2.4,13a) we get the trimmed lift curve:

CmoCL6• detCLmm = - + -- am

Cm6• <; ~ (2.4,15)

and the slope is given by

tCL) CL6-- =C ----=-Cda trim L", Cm6• m.,

The trimmed lift-curve slope is seen to be less than CL'"by an amount that depends onCm"" i.e., on the static margin, and that vanishes when h = hn• The difference is onlya few percent for tailed airplanes at normal CG position, but may be appreciable fortailless vehicles because of their larger CL6' The relation between the basic andtrimmed lift curves is shown in Fig. 2.17.

Equation (2.4,13b) is plotted on Fig. 2.18, showing how 5etrim varies with CLmmand CG position when the aerodynamic coefficients are constant.

(2.4,16)

Page 47: Stability Control EtkinReid

2.4 Longitudinal Control 37

Basic(6.- 0)II

/ Trimmed/

I.

~~f=------------....•a

Figure 2.17 Trimmed lift curve.

VARIATION OF Betrlm WITH SPEEDWhen, in the absence of compressibility, aeroelastic effects, and propulsive systemeffects, the aerodynamic coefficients of (2.4,13) are constant, the variation of 8etrim

with speed is simple. Then 8etrim is a unique function of CLmm for each CG position.Since CLmm is in tum fixed by the equivalent airspeed," for horizontal flight

wCLrrim = 1 V 2S

2PO E(2.4,17)

then Setrim becomes a unique function of VE• The form of the curves is shown in Fig.2.19 for representative values of the coefficients.

The variation of Setrim with CLrrim or speed shown on Figs. 2.18 and 2.19 is thenormal and desirable one. For any CG position, an increase in trim speed from anyinitial value to a larger one requires a downward deflection of the elevator (a forward

Figure 2.18 Elevator angle to trim at various CG positions.

4Equivalent airspeed (EAS) is VE = V~ where Po is standard sea-level density.

Page 48: Stability Control EtkinReid

38 Chapter 2. Static Stability and Control-Part 1

Figure 2.19 Example of variation of elevator angle to trim with speed and CG position.

movement of the pilot's control). The "gradient" of the movement o5elrirliJVE is seento decrease with rearward movement of the CO until it vanishes altogether at the NP.In this condition the pilot in effect has no control over trim speed, and control of thevehicle becomes very difficult. For even more rearward positions of the CO the gra-dient reverses, and the controllability deteriorates still further.

When the aerodynamic coefficients vary with speed, the above simple analysismust be extended. In order to be still more general, we shall in the following explic-itly include propulsive effects as well, by means of the parameter 8p, which stands forthe state of the pilot's propulsion control (e.g., throttle position). 5p = constant there-fore denotes fixed-throttle and, of course, for horizontal flight at varying speed, 5p

must be a function of V that is compatible with T = D. For angles of climb or descentin the normal range of conventional airplanes L == W is a reasonable approximation,and we adopt it in the following. When nonhorizontal flight is thus included, 8p be-comes an independent variable, with the angle of climb 'Y then becoming a functionof 5p, V, and altitude.

The two basic conditions then, for trimmed steady flight on a straight line are

Cm=OL = CLtpV2S = W

(2.4,18)

and in accordance with the postulates made above, we write

c; = c.i« V, s; 5p)

CL = CL(a, V, s; 5p)

Now let ( )e denote one state that satisfies (2,4,18) and consider a small changefrom it, denoted by differentials, to another such state. From (2.4,18) we get, for p =constant,

(2.4,19)

anddCm=O

CLV2 = const2VeCLedV + V;dCL = 0

dV A

dC = -2C - = -2C dVL LeV L.e

(2.4,20)

or

so that (2.4,21)

Page 49: Stability Control EtkinReid

2.4 Longitudinal Control 39

where V is VIVe• Taking the differentials of (2.4,19) and equating to (2.4,20 and 21)we get

CLada + CLa,d5e = -CLa1'd5p - (CLv + 2C1.)dVCmada + Cma,d5e = -Cmapd5p - CmvdV

where CLv = oCJoVand Cmv = oC,joV. From (2.4,22) we get the solution for d5e as

as, = d:t {[(CLV + 2CLe)Cma - CLaCmv]dV + (CLapCma- CLaCmap)d8p} (2.4,23)

There are two possibilities, 8p constant and 8p variable. In the first case (fixed throt-tle), d8p = 0 and

(2.4,22)

(CLV + 2CL)Cma - CLaCmvdet

(2.4,24)

It will be shown in Chap. 6 that the vanishing of this quantity is a true criterion ofstability, that is it must be >0 for a stable airplane. In the second case, for exampleexactly horizontal flight, 5p = 5/V) and the 5p term on the right-hand side of (2.4,23)remains. For such cases the gradient (d5elrijdV) is not necessarily related to stability.For purposes of calculating the propulsion contributions, the terms CLa d5p and

p

Cmapd5p in (2.4,23) would be evaluated as dCLp and dCmp [see the notation of (2.3,1)].These contributions to the lift and moment are discussed in Sec. 3.4.

The derivatives CLvand Cmv may be quite large owing to slipstream effects onSTOL airplanes, aeroelastic effects, or Mach number effects near transonic speeds.These variations with M can result in reversal of the slope of 5elrim as illustrated onFig. 2.20. The negative slope at A, according to the stability criterion referred toabove, indicates that the airplane is unstable at A. This can be seen as follows. Let theairplane be in equilibrium flight at the point A, and be subsequently perturbed so thatits speed increases to that of B with no change in a or 5e• Now at B the elevator angleis too positive for trim: that is there is an unbalanced nose-down moment on the air-plane. This puts the airplane into a dive and increases its speed still further. The speedwill continue to increase until point C is reached, when the Be is again the correctvalue for trim, but here the slope is positive and there is no tendency for the speed tochange any further.

Ol--------~.,."e:.:..------------_v

Figure 2.20 Reversal of B'trim slope at transonic speeds, B1'= const.

Page 50: Stability Control EtkinReid

40 Chapter 2. Static Stability and Control-Part 1

STATIC STABILITY LIMIT, h.

The critical CG position for zero elevator trim slope (i.e, for stability) can be foundby setting (2.4,24) equal to zero. Recalling that Cma = CLJh - hn), this yields

(2.4,25)

or

c;where h, = hn + (2.4,26)

CLy + 2CLe

Depending on the sign of Cmv' h, may be greater or less than hn' In terms of hs'(2.4,24) can be rewritten as

(d5etrim) CLa~ B

p= det (CLy + 2CL)(h - hs)

(h - hs) is the "stability margin," which may be greater or less than the static margin.

(2.4,27)

FLIGHT DETERMINATION OF hn AND h.

For the general case, (2.3,5) suggests that the measurement of hn requires the mea-surement of Cma and CLa' Flight measurements of aerodynamic derivatives such asthese can be made by dynamic techniques. However, in the simpler case when thecomplications presented by propulsive, compressibility, or aeroelastic effects are ab-sent, then the relations implicit in Figs. 2.18 and 2.19 lead to a means of finding h;from the elevator trim curves. In that case all the coefficients of (2.4,13) are con-stants, and

(2.4,28)dCLmm det

d5etrim = _ CLa (h - h )dCLmm det n

Thus measurements of the slope of 5etrim vs. CLmm at various CG positions produce acurve like that of Fig. 2.21, in which the intercept on the h axis is the required NP.

When speed effects are present, it is clear from (2.4,27) that a plot of (d5etrimldV)8pagainst h will determine h, as the point where the curve crosses the h axis.

or (2.4,29)

Figure 2.21 Determination of stick-fixed neutral point from flight test.

Page 51: Stability Control EtkinReid

2.5 The Control Hinge Moment 41

2.5 The Control Hinge MomentTo rotate any of the aerodynamic control surfaces, elevator, aileron, or rudder, aboutits hinge, it is necessary to apply a force to it to overcome the aerodynamic pressuresthat resist the motion. This force may be supplied entirely by a human pilot through amechanical system of cables, pulleys, rods, and levers; it may be provided partly by apowered actuator; or the pilot may be altogether mechanically disconnected from thecontrol surface ("fly-by-wire" or "fly-by-light"). In any case, the force that has to beapplied to the control surface must be known with precision if the control system thatconnects the primary controls in the cockpit to the aerodynamic surface is to be de-signed correctly. The range of control system options is so great that it is not feasiblein this text to present a comprehensive coverage of them. We have therefore limitedourselves in this and the following chapter to some material related to elevator con-trol forces when the human pilot supplies all of the actuation, or when a power assistrelieves the pilot of a fixed fraction of the force required. This treatment necessarilybegins with a discussion of the aerodynamics; that is, of the aerodynamic hinge mo-ment.

The aerodynamic forces on any control surface produce a moment about thehinge. Figure 2.22 shows a typical tail surface incorporating an elevator with a tab.The tab usually exerts a negligible effect on the lift of the aerodynamic surface towhich it is attached, although its influence on the hinge moment is large.

Hingeline

Elevator---r---_T - _

~ -Tab-I ~ Tab hinge

~se_A ~1(a)

toE--------c,------(b)

Figure 2.22 Elevator and tab geometry. (a) Plan view. (b) SectionA-A.

Page 52: Stability Control EtkinReid

42 Chapter 2. Static Stability and Control-Part 1

The coefficient of elevator hinge moment is defined by

HeCh =---

e ipy2sece

Here He is the moment, about the elevator hinge line, of the aerodynamic forces onthe elevator and tab, Se is the area of that portion of the elevator and tab that lies aft ofthe elevator hinge line, and ce is a mean chord of the same portion of the elevator andtab. Sometimes ce is taken to be the geometric mean value, that is, ce = Sj2se, andother times it is the root-mean square of Ceo The taper of elevators is usually slight,and the difference between the two values is generally small. The reader is cautionedto note which definition is employed when using reports on experimental measure-ments of Che•

Of all the aerodynamic parameters required in stability and control analysis, thehinge-moment coefficients are most difficult to determine with precision. A large.number of geometrical parameters influence these coefficients, and the range of de-sign configurations is wide. Scale effects tend to be larger than for many other pa-rameters, owing to the sensitivity of the hinge moment to the state of the boundarylayer at the trailing edge. Two-dimensional airfoil theory shows that the hinge mo-ment of simple flap controls is linear with angle of attack and control angle in bothsubsonic and supersonic flow.

The normal-force distributions typical of subsonic flow associated with changesin a and Se are shown qualitatively in Fig. 2.23. The force acting on the movable flaphas a moment about the hinge that is quite sensitive to its location. Ordinarily thehinge moments in both cases (a) and (b) shown are negative.

In many practical cases it is a satisfactory engineering approximation to assumethat for finite surfaces Che is a linear function of as, Se' and St. The reader shouldnote however that there are important exceptions in which strong nonlinearities arepresent.

We assume therefore that Che is linear, as follows,

Che = bo + bias + bzl>e + b3l>t (2.5,1)

ec;where bI = -::1- = Cheoa

sas

sc;bz = ij"5 = CheSe

e

aChe

b3 = aa- = CheStt

as is the angle of attack of the surface to which the control is attached (wing or tail),and 81is the angle of deflection of the tab (positive down). The determination of thehinge moment then resolves itself into the determination of bo, b., bz, and b3• Thegeometrical variables that enter are elevator chord ratio cfc; balance ratio cJce, noseshape, hinge location, gap, trailing-edge angle, and planform. When a set-back hingeis used, some of the pressure acts ahead of the hinge, and the hinge moment is lessthan that of a simple flap with a hinge at its leading edge. The force that the controlsystem must exert to hold the elevator at the desired angle is in direct proportion tothe hinge moment.

Page 53: Stability Control EtkinReid

2.5 The Control Hinge Moment 43

:J---~

Fixed surface

(al

V'

Figure 2.23 ·Nonnal-force distribution over control surface at subsonic speed. (a) Forcedistribution over control associated with at at 8. = O. (b) Force distribution over control associatedwith 8. at zero at.

We shall find it convenient subsequently to have an equation like (2.5,1) with ainstead of as. For tailless aircraft, as is equal to a, but for aircraft with tails, as = at,Let us write for both types

(2.5,2)

where for tailless aircraft Ch•o = bo, Che", = bI, For aircraft with tails, the relation be-tween a and at is derived from (2.3,12) and (2.3,19), that is,

(2.5,3)

whence it follows that for tailed aircraft, with symmetrical airfoil sections in the tail,for which bo = 0,

(a)(2.5,4)

(b)

Page 54: Stability Control EtkinReid

44 Chapter 2. Static Stability and Control-Part 1

2.6 Influence of a Free Elevator on Lift and MomentIn Sec. 2..3 we have dealt with the pitch stiffness of an airplane the controls of whichare fixed in position. Even with a completely rigid structure, which never exists, amanually operated control cannot be regarded as fixed. A human pilot is incapable ofsupplying an ideal rigid constraint. When irreversible power controls are fitted, how-ever, the stick-fixed condition is closely approximated. A characteristic of interestfrom the point of view of handling qualities is the stability of the airplane when theelevator is completely free to rotate about its hinge under the influence of the aerody-namic pressures that act upon it. Normally, the stability in the control-free conditionis less than with fixed controls. It is desirable that this difference should be small.Since friction is always present in the control system, the free control is never real-ized in practice either. However, the two ideal conditions, free control and fixed con-trol, represent the possible extremes.

When the control is free, then Che = 0, so that from (2.5,2)

I5efree = - b

2(Cheo + Che,p. + b30t) (2.6,1)

The typical upward deflection of a free-elevator on a tail is shown in Fig. 2.24. Thecorresponding lift and moment are

..

CLrree = CLaa + CLaeoe_

Cmtree = c.; + Cmaa. + CmlJe5eftw

After substituting (2.6,1) into (2.6,2), we get

CLrree = C~ + CLa' a

Cmftw = C:"o + C;"aa

(2.6,2)

(a)(2.6,3)

(b)

CC~ = - ~lJe (Cheo + b30t)

r - C' - C _ CLaeCheaa - L - La a b

2

CC:..o = Cmo - b:ae (Cheo + b35t)

C CC' = C _ ma. hearna ma b

2

When due consideration is given to the usual signs of the coefficients in these equa-tions, we see that the two important gradients CLa and Cma are reduced in absolute

where (a)(2.6,4)

(b)

(a)(2.6,5)

(b)

Figure 2.24 Elevator floating angle.

Page 55: Stability Control EtkinReid

2.6 Influence of a Free Elevator on Lift and Moment 45

magnitude when the control is released. This leads, broadly speaking, to a reductionof stability.

FREE-ELEVATOR FACTORFor a tailless aircraft with a free elevator, the lift-curve slope is (cf 2.6,4b)

a' = a(1 - CL8}l) (2.6,6)ab2

The factor in parentheses is the free elevator factor, and normally has a value lessthan unity. When the elevator is part of the tail, the floating angle can be related to at,viz forbo = 0

Che = blat + b28efree + b38t = 0

1or 8efree = - b

2(bl at + b38t)

and the tail lift coefficient is

(2.6,7)

(2.6,8)

The effective lift-curve slope is

dC~,-:1- = Fat (2.6,9)oat

where F = (I - ae !!..:...) is the free elevator factor for a tail. If Fat is used in placeat b2

of at and a' in place of a, then all the equations given in Sec. 2.3 hold for aircraftwith a free elevator.

ELEVATOR·FREE NEUTRAL POINT

It is evident from the preceding comment that the NP of a tailed aircraft when the el-evator is free is given by (2.3,23) as

Fat _ ( dE ) 1 ec;h~ = h; b + -, VH 1 - ~ - -, ~

w a oa a oa

Alternatively, we can derive the NP location from (2.6,5b), for we know from (2.3,5)that

(2.6,10)

or

(a)

(2.6,11)

(b)

Page 56: Stability Control EtkinReid

46 Chapter 2. Static Stability and Control-Part 1

Since CmBe is of different form for the two main types of aircraft, we proceed sepa-rately below.

TAILED AIRCRAFT

CmBe is given by (2.4,8), so (2.6,11) becomes for this case

Using (2.6,4b) this becomes

(2.6,12)

Finally, using (2.4,8) for CLBe, and (2.5,4) for c.; we get

a; b, ( oe ) ( 8, _)h' = h - - - 1 - - -(h - h ) - + Vn n a' b

2ocr n nwb 8 H

(2.6,13)

TAILLESS AIRCRAFT

CmBe is given by (2.4,9) and Che", = b., When these are substituted into (2.6,11) the re-sult is

By virtue of (2.6,6) this becomes

or

b oCh - h' = h - h __ 1_ ~

n n a'b2 oBe

b oCh' =h + _1_ ~n n a'b2 oBe

(2.6,14)

The difference (h~ - h) is called the control-free static margin, K~.When representa-tive numerical values are used in (2.6,13) one finds that h; - h~may be typicallyabout 0.08. This represents a substantial forward movement of the NP, with conse-quent reduction of static margin, pitch stiffness, and stability.

Page 57: Stability Control EtkinReid

2.7 The Useo/Tabs 47

2.7 The Use of TabsTRIM TABS

In order to fly at a given speed, or CL, it has been shown in Sec. 2.4 that a certain ele-vator angle 8etrlm is required. When this differs from the free-floating angle 8err.e' aforce is required to hold the elevator. When flying for long periods at a constantspeed, it is very fatiguing for the pilot to maintain such a force. The trim tabs areused to relieve the pilot of this load by causing 8etrlm and 8err.e to coincide. The trim-tab angle required is calculated below.

When Che and Cm are both zero, the tab angle is obtained from (2.5,2) as

(2.7,1)

On substituting from (2.4,13) (which implies neglecting oCm/o8t), we get

1 [Cmo CLmm ]8'trlm = - b

3Cheo + det (CheaCLS• - b2CL) + det (Che"Cms. - b2Cm)

which is linear in CLmm for constant h, as shown in Fig. 2.25. The dependence on h issimple, since from (2.6,11) we find that

(CheaCms. - b2Cm) = -a'b2(h - h~)

and hence

1 [Cmo a'b2]8'trlm = - b

3Cheo + det (Che"CLs• - b2CL) - det (h - h~)CLmm (2.7,2)

This result applies to both tailed and tailless aircraft, provided only that the appropri-ate values of the coefficients are used. It should be realized, of course, in reference toFig. 2.25, that each different CLmm in a real flight situation corresponds to a differentset of values of M, tpV:z, and CT> so that in general the coefficients of (2.7,2) varywith CL> and the graphs will depart from straight lines.

--::o:+---...,~.....-=::::;....----~ CLtrlm

Figure 2.25 Tab angle to trim.

Page 58: Stability Control EtkinReid

48 ChIlpter2. Static Stability and Control-Part 1

Equation (2.7,2) shows that the slope of the 8ttrimvs CLtrim curve is proportional tothe control-free static margin. When the coefficients are constants, we have

d8ttrim b2 a'-dC = -b -d (h- h~) (2.7,3)

Ltrim 3 et

The similarity between (2.7,3) and (2.4,13c) is noteworthy, that is the trim-tab slopebears the same relation to the control-free NP as the elevator angle slope does to thecontrol-fixed NP. It follows that flight determination of h~ from measurements ofd8ttrim/dCLtrim is possible subject to the same restrictions as discussed in relation to themeasurement of hn in Sec. 2.4.

OTHER USES

Tabs are used for purposes other than trimming, especially for manually actuatedcontrols. Three of the main types are as follows:

Geared Tabs. Tabs connected to the main surface by a mechanical linkage thatcauses the tab to deflect automatically when the main surface is deflected, butin the opposite direction. The hinge moment produced by the tab then assiststhe rotation of the main surface. These have the effect of reducing the b2 of thesurface.

Spring Tabs. Tabs connected to the main surface by an elastic element. Thedesign is such that the deflection of the tab depends on the dynamic pressurein a way that mitigates the effect of the speed-squared law on the control force.

Servo Tabs. An arrangement in which the pilot controls the tab directlythrough a mechanical linkage. It is then the tab, not the pilot, that provides thehinge moment needed to rotate the main surface.

••Both spring tabs and servo tabs are effective devices for reducing control forces

on large high-speed airplanes. However, both add an additional degree of freedom tothe control system dynamics, and this is a potential source of trouble due to vibrationor flutter.

For further details of how these tabs function, see Etkin (1972).

2.8 Control Force to TrimThe importance of control forces in relation to handling qualities has already beenemphasized in previous sections, and the many options available to designers of pow-ered control systems has been noted. Cockpit devices can of course be designed toproduce more or less any desired synthetic feel on the primary flight controls. It isnevertheless both instructive and necessary to be able to calculate the control forcesthat will be present in the case of natural feel, or when a simple power assist is pres-ent. A case in point is the elevator force required to trim the airplane, and how itvaries with flight speed.

Figure 2.26 is a schematic representation of a reversible control system. The boxdenoted "control system linkage" represents any assemblage of levers, rods, pulleys,

Page 59: Stability Control EtkinReid

2.8 Control Force to Trim 49

n1+-.-.-p,.

--r-E--~e----1'-Contro,-syste;:-lf-.---..:::::====

I linkage I'-------_---1Figure 2.26 Schematic diagram of an elevator control system.

cables, and power-boost elements that comprise a general control system. We assumethat the elements of the linkage and the structure to which it is attached are ideallyrigid, so that no strain energy is stored in them, and we neglect friction. The systemthen has one degree of freedom. P is the force applied by the pilot, (positive to therear) s is the displacement of the hand grip, and the work done by the power boostsystem is Wb• Considering a small quasistatic displacement from equilibrium (i.e., nokinetic energy appears in the control system), conservation of energy gives

Pds + dWb + Hed8e = 0dWb as,

P=----Hds ds e

Now the nature of ratio or power boost controls is such that dW z/ds is proportional toP or He. Hence we can write

(2.8,1)

or

(2.8,2)

and

as,GI = - ds > 0, the elevator gearing (rad/ft or rad/m)

dWJdsG2 = H ' the boost gearing (ft-I or m")

e

where

(2.8,2) is now rewritten as

(2.8,3)

where G = GI - G2. For fixed GI, i.e., for a given movement of the control surfaceto result from a given displacement of the pilot's control, then the introduction ofpower boost is seen to reduce G and hence P. G may be designed to be constant overthe whole range of 8e, or it may, by the use of special linkages and power systems, bemade variable in almost any desired manner.

Introduction of the hinge-moment coefficient gives the expression for P as

P = GCheSeC/~pV2 (2.8,4)

and the variation of P with flight speed depends on both V2 and on how Che varieswith speed.

Page 60: Stability Control EtkinReid

50 Chapter 2. Static Stability and Control-Part 1

The value of Che at trim for arbitrary tab angle is given by

Che = Cheo + Che"Umm + bzl)etrim+ b3l)t (2.8,5)

(2.8,5) in combination with (2.7,1) yields

c; = bil)t - l)ttrim) (2.8,6)

i.e., the hinge moment is zero when l)t = l)ttrim as expected, and linearly proportionalto the difference. From (2.7,2) then the hinge moment is

C~ a'bzChe = b38t + Cheo + - (Che CLB - bZCL ) - -----;,- (h - h~)C. (2.8,7)det ". "uet e-mm

Lift equals the weight in horizontal flight, so that

W

CLmm = tpVZ

where w = W/S is the wing loading. When (2.8,7 and 8) are substituted into (2.8,4)the result obtained is

where

(2.8,8)

P =A + BtpVZ (2.8,9)

a'bA = -GSi5ew ~ (h - h~)

aet

_ [ C~ ]B = GSece b38t + Cheo + det (Che"CLlie - bZCL)

The typical parabolic variation of P with V when the aerodynamic coefficientsare all constant, is shown in Fig. 2.27. The following conclusions may be drawn.

p

..•

Const term

',>0Figure 2.27 Example of low-speed control force.

Page 61: Stability Control EtkinReid

2.9 Control Force Gradient 51

1. Other things remaining equal. P oc Sece. Le., to the cube of the airplane size.This indicates a very rapid increase in control force with size.

2. P is directly proportional to the gearing G.3. The CO position only affects the constant term (apart from a second-order in-

fluence on em,;)' A forward movement of the CO produces an upward transla-tion of the curve.

4. The weight of the airplane enters only through the wing loading, a quantitythat tends to be constant for airplanes serving a given function, regardless ofweight. An increase in wing loading has the same effect as a forward shift ofthe CO.

5. The part of P that varies with ipV2 decreases with height, and increases as thespeed squared.

6. Of the terms contained in E, none can be said in general to be negligible. Allof them are "built-in" constants except for 5,.

7. The effect of the trim tab is to change the coefficient of ipV2, and hence the

curvature of the parabola in Fig. 2.27. Thus it controls the intercept of thecurve with the V axis. This intercept is denoted Vtrim; it is the speed for zerocontrol force .

•2.9 Control Force Gradient

It was pointed out in Sec. 2.7 how the trim tabs can be used to reduce the controlforce to zero. A significant handling characteristic is the gradient of P with Vat P =O. The manner in which this changes as the CO is moved aft is illustrated in Fig.2.28. The trim tab is assumed to be set so as to keep Vtrim the same. The gradientdPldV is seen to decrease in magnitude as the CO moves backward. When it is at thecontrol-free neutral point, A = 0 for aircraft with or without tails, and, under thestated conditions, the PIV graph becomes a straight line lying on the V axis. This isan important characteristic of the control-free NP; that is, when the CO is at thatpoint, no force is required to change the trim speed.

A quantitative analysis of the control-force gradient follows.

p

-L.-I,L-I-------~:a...,,....---------vA~A>hl

--- ...••

Figure 2.28 Effect of CG location on control-force gradient at fixed trim speed.

Page 62: Stability Control EtkinReid

52 Chapter 2. Static Stability and Control-s-Part 1

The force is given by (2.8,9). From it we obtain the derivative

CJPCJV = BpV

At the speed Vtrim, P = 0, and B = -A1ipV2trim' whence

CJP 2A-=---CJV Vtrim

A is given following (2.8,9). Substituting the value into (2.9,1) we get

CJP a'b2 w- = 2GSc - - (h -h')CJV e e det Vtrim n

From (2.9,2) we deduce the following:

1. The control-force gradient is proportional to Sece; that is, to the cube of air-plane size.

2. It is inversely proportional to the trim speed; i.e. it increases with decreasingspeed. This effect is also evident in Fig. 2.27.

3. It is directly proportional to wing loading.4. It is independent of height for a given true airspeed, but decreases with height

for a fixed VE•

5. It is directly proportional to the control-free static margin.

(2.9,1)

(2.9,2)

Thus, in the absence of Mach number effects, the elevator control will be "heaviest"at sea-level, low-speed, forward CG, and maximum weight.

2.10 Exercises2.1 A subsonic transport aircraft has a tapered, untwisted sweptback wing with straight

leading and trailing edges. The wing tips are straight and parallel to the root chord. Inthe following, use the data of Appendix C and assume that the airfoil section localaerodynamic center is at the ~-chord point.

(a) Make an accurate three-view drawing of the wing chord plane.(b) Calculate wing area S, aspect ratio A, taper ratio A = c,lcr and the mean aerody-

namic chord c.(c) Calculate the location of the wing's mean aerodynamic center, and locate it and c

on the side view of the wing (with dimensions). (Assume a uniform additionallift coefficient Cia = CL.)

(d) The aircraft is to be operated with its most rearward CG position limited to 25 ft(7.62 m) aft of the apex of the wing. The distance between the wing and tailmean aerodynamic centers is Zt :::: 55 ft (16.76 m), Estimate the tail area requiredto provide a control-fixed static margin of at least 0.05 at all times. Assume thatat = awb and hnw = hnwo' Ignore power plant effects and use CJe/CJa= 0.25.

Geometric Data

..

Wing Span, b

Root Chord, c,Tip Chord, c,

150 ft (45.72 m)

25 ft (7.62 m)12 ft (3.66 m)

Page 63: Stability Control EtkinReid

2.10 Exercises 53

Leading edge sweep, Ao 26°Dihedral angle, 'Y 4°

2.2 Evaluate the validity of the approximation made in going from (2.2,2) to (2.2,3) byusing the data for the airplane of Exercise (2.1) and calculating Cmw from both equa-tions. Assume CD = CD. + KCL and L = Lw• The following additional data areprovided. w OllOW

Geometric DataWeight, Wzlc

207,750lb (924,488 N)

0.15

Aerodynamic Dataaw 0.080/deg

-0.050.0130.054350 kts (180 mls)

2.377 X 10-3 slugs/ft" (1.225 kg/m")

KVp

2.3 Show that if CmWb is a linear function of CLwb then Cmacwb is a constant.

2.4 Beginning with (2.3,1) perform the reductions to derive (2.3,20) to (2.3,23).

2.5 The following data apply to a i'5 scale wind tunnel model of a transport airplane. Thefull-scale mass of the aircraft is 1,552.80 slugs (22,680 kg). Assume that the aerody-namic data can be applied at full-scale. For level unaccelerated flight at V = 239knots (123 mls) of the full-scale aircraft, under the assumption that propulsion effectscan be ignored,

(a) Find the limits on tail angle it and CG position h imposed by the conditionsc.; > 0 and Cma < O.

(b) For trimmed flight with 5e = 0, plot it vs. h for the aircraft and indicate wherethis line meets the boundaries of part (a).

Geometric DataWing area, SWing mean aerodynamic chord, cItTail area, s,

1.50 :ff (0.139 m2)

6.145 in (15.61 em)15.29 in (38.84 em)0.368 ft2 (0.0342 m2

)

Aerodynamic Dataawb 0.077/degat 0.064/degEo 0.72°

0.30

-0.018

Page 64: Stability Control EtkinReid

54 Chapter 2. Static Stability and Control-s-Par: 1

.50

.25

t.J~

-.25

-.50

-.75

Figure 2.29 Data for Exercise 2.6.

«: 0.25p 2.377 X 10-3 slugs/ft" (1.225 kg/m")

2.6* The McDonnell Douglas C-l7 is a four-engined jet STOL transport airplane.

(a) Find A and c for the wing using the geometrical data and Appendix C.(b) Use Appendix B to estimate aw' the wing lift curve slope, assuming that f3 = 1

and K = 1.(c) If at = 0.068/deg and awb = aw, find the lift curve slope, a, of the aircraft. As-

de 2awsume da = TTA (with aw expressed in rad ").

(d) Find ema for the case where It = It = 92 ft (28.04 m). Ignore propulsion effects.

*Problem courtesy of Professor E. K. Parks, University of Arizona.

10°

0

~_5°

_10°

_150

••

Figure 2.30 Trim data for Exercise 2.6.

Page 65: Stability Control EtkinReid

2.10 Exercises 55

(e) From the experimental curves of Figs. 2.29 and 2.30 and the given geometry, findCmBe and hn- Find c.; for h = 0.30.

Geometric Data

Wing area, SWing span, b

Root chord, cr

Tip chord, c,ichord line sweep, Aichord line sweep, Acl2

Tail area, s,

3,800 fe (353.0 rrr')

165 ft (50.29 m)37.3 ft (11.37 m)

8.8 ft (2.68 m)

25°22°870 fe (80.83 rrr')

2.7 Consider an aircraft with its tail identical to its wing (i.e., the same span, area, chord,etc.). Neglect body and wing-body interaction effects [i.e., in general ( )wb ==( )wJ, neglect propulsion effects and assume zero elevator and tab deflections. As-sume (2.2,7) in this instance is approximated by M, = -l,Lt + Mac,.

(a) What changes should be made in the expressions for a (2.3,18), Cma (2.3,21a),and Cmo (2.3,22a)?

aE(b) What would aa have to be numerically in order that the neutral point hn lies

midway between the mean aerodynamic centers of the wing and tail?(c) For trimmed level flight, derive an expression for the ratio of the lift generated by

the wing to the lift generated by the tail as a function of the tail angle it. AssumeaEEo = 0, -a = 0.2, a = 5 (rad"), Cm = 0.2 and (h - hn) = -0.3.a 0

2.8* The following data were taken from a flight test of a PA-32R-300 Cherokee-6 air-plane.

Altitude VE Mass i, XCG

(ft) (km) (mph) (m/s) (slugs) (kg) (deg) (in) (em)

4540 1.384 91.0 40.7 113.4 1656 1.5 93.89 238.54560 1.390 109 48.7 113.0 1650 0 93.89 238.54700 1.433 126 56.3 112.9 1649 -1.0 93.89 238.54580 1.396 155 69.3 112.7 1646 -2.0 93.89 238.55320 1.622 89.0 39.8 100.4 1466 4.5 86.82 220.54620 1.408 105 46.9 100.2 1463 2.0 86.82 220.54740 1.445 123 55.0 100.0 1461 0.3 86.82 220.54900 1.494 151 67.5 99.84 1458 -1.0 86.82 220.54880 1.487 87.0 38.9 88.51 1293 7.2 80.43 204.34820 1.469 103 46.0 88.35 1290 3.5 80.43 204.34880 1.487 122 54.5 88.20 1288 1.5 80.43 204.34740 1.445 152 68.0 88.04 1286 0 80.43 204.3

*Problem courtesy of Professor E. K. Parks, University of Arizona.

Page 66: Stability Control EtkinReid

56 Chapter 2. Statk Stability and Control-Part 1

The data were taken in trimmed level flight. XCG is the distance of the CG aft of thenose of the aircraft. The aircraft has an all-moving tail and thus it is used instead of Sto trim the aircraft. The wing area is S = 174.5 ff (16.21 m2). e

(a) Plot tail-setting angle, it, versus the lift coefficient of the aircraft for each of thethree CG locations.

(b) Curve fit the data points in (a) with three straight lines having a common inter-cept (refer to Fig. 2.18).

(c) Use a graphical technique to find the location of the neutral point (controls fixed)relative to the nose of the aircraft (refer to Fig. 2.21).

2.9 Starting with (2.6,llb), derive (2.6,13).

2.10 The elevator control force to trim a particular airplane at a speed of 300 kts (154 mls)is zero. Using the following data estimate the force required to change the trim speedto 310 kts (159 mls). Assume that CLBe is sufficiently small that CLBe = 0 can be usedin the expression for control force.

Geometric DataElevator gearing, G

Elevator area aft of hinge line, S,Mean elevator chord, ce

VH

CG location, hWing loading, w

Aerodynamic DataElevator hinge moment coefficient,sc;aSe

ae

Neutral point, elevator free, h~

3°/in (1.18°/cm)40 ft2 (3.72 m2)2.0 ft (0.61 m)0.560.3850 psf (2,395 Pa)

-0.OO5/deg

0.025/deg0.45

2.11 A fatal airplane accident has led to a civil court case. It is alleged that the airplane inquestion was unstable-"that the neutral point was ahead of the center of gravity."

You are called as an expert witness to explain to the court (i.e., to the judge and theattorneys) the meaning and importance of stability, center of gravity, and neutralpoint. You know that your audience is composed of intelligent persons educated inthe humanities and law, but you must assume that they have only a rudimentaryknowledge of science and mathematics.

Write an essay describing how you would meet this challenge. You are free to use di-agrams and simple models, but you must avoid any use of mathematics. You shouldkeep your language simple, avoiding any technical jargon. Even the word momentshould be avoided. Your goal is to clarify, not to impress the court with your superiortechnical knowledge.

Page 67: Stability Control EtkinReid

a= CLOI

a' = C'LOI

atbo = Cheo

bi = CheOl'

b2 = CheSe

b3 = CheStC

CDc:CLc.:CL,

CLSe

CLv

Cm

c;Cmo

c.;c;Cmv

c;c.;CTdet

2.11 Additional Symbols Introduced in Chapter 2 57

2.11 Additional Symbols Introduced in Chapter 2

airplane lift-curve slope, ()CL/()a, elevator fixedairplane lift-curve slope, 'dC~/'da, elevator free'dCL/'d5e

wing lift-curve slope, 'dCLw/'dawing-body lift-curve slope, 'dCLw/'datail lift-curve slope, 'dCL/'dat

see Eq. 2.5,1'dCh)'dat

'dCh)'d5e

'dCh)'d5t

length of mean aerodynamic chordmean elevator chord (see Sec. 2.5)D/!plftSelevator hinge-moment coefficient, HAplftseceu~plfts

LwJiplftstail lift coefficient, L/iplftst

'dCL/'d5e

aCd'dVM/~plftSc

MacAplftScMacjiplftScairplane pitching-moment coefficient at zero aairplane pitching-moment coefficient at zero awb

pitching-moment coefficient of the propulsion unitsM/iplftSc'dCmI'dV'dCmI'da'dCmI'd5e

T/iplftssee (2.4,14)wing dragdrag of the tailfree-elevator factor (l - aebrlaP2) for aircraft with tailselevator gearing

elevator hinge moment

Page 68: Stability Control EtkinReid

58 Chapter 2. Static Stability and Control-Part 1

h CG position, fraction of mean chord (see Fig. 2.9)hn neutral point of airplane, fraction of mean chord, elevator fixedh' neutral point of airplane, fraction of mean chord, elevator freen

hnw neutral point of wing, fraction of mean chordhnWb neutral point of the wing-body combinationhs static stability limit

it tail-setting angle (see Fig. 2.11)s, static margin, see (2.3,6)L airplane lift

Lw wing liftLWb lift of wing-body combinationLt lift of the tail

t, distance between CG and tail mean aerodynamic centert, see Fig. 2.12M Mach numberM pitching moment about the CGMw pitching moment of the wing about the CGMacw pitching moment of the wing about its mean aerodynamic centerMacWb pitching moment of the wing-body combination about its mean ..

aerodynamic centerMWb pitching moment, about the CG, of the wing-body combination

Mt pitching moment of tail about CG

P control force, positive to the rear

S wing area

s, span of elevator

Se area of elevator aft of hinge line

s, area of tail

T thrust

V true airspeed

V VIVeVe reference equilibrium airspeed

VE equivalent airspeed (EAS), V~

VB horizontal tail volume, Stl/Se

VB Sl/Se

W aircraft weight

w wing loading (W/S)

a angle of attack of the zero lift line of the airplane (elevator anglezero)

"'.

Page 69: Stability Control EtkinReid

2.11 Additional Symbols Introduced in Chapter 2 59

angle of attack of the zero lift line of the wingangle of attack of the zero lift line of the wing-body combinationangle of attack of the tailangle of attack of the thrust lineelevator angletab angledownwash angledownwash when fXwb = 0air densitystandard sea-level value of p (see Appendix D)

PPo

Page 70: Stability Control EtkinReid

CHAPTER 3

Static Stability and Control-Part 2

3.1 Maneuverability-Elevator Angle per gIn this and the following sections, we investigate the elevator angle and control forcerequired to hold the airplane in a steady pull-up with load factor' n (Fig. 3.1). Theconcepts discussed here were introduced by S. B. Gates (1942). The flight-path tan-gent is horizontal at the point under analysis, and hence the net normal force is L -W = (n - I)Wvertically upward. The normal acceleration is therefore (n - l)g.

When the airplane is in straight horizontal flight at the same speed and altitude,the elevator angle and control force to trim are Se and P, respectively. When in thepull-up, these are changed to Be + liBe and P + liP. The ratios lifJ)(n - 1) andliPI(n - 1) are known, respectively, as the elevator angle per g, and the control forceper g. These two quantities provide a measure of the maneuverability of the airplane;the smaller they are, the more maneuverable it is.

The angular velocity of the airplane is fixed by the speed and normal accelera-tion (Fig. 3.1).

q=(n - l)g

V(3.1,1)

As a consequence of this angular velocity, the field of the relative air flow past theairplane is curved. It is as though the aircraft were attached to the end of a whirlingarm pivoted at 0 (Fig. 3.1). This curvature of the flow field alters the pressure distri-bution and the aerodynamic forces from their values in translational flight. Thechange is large enough that it must be taken into account in the equations describingthe motion.

We assume that q and the increments Aa, ABe etc. between the rectilinear andcurved flight conditions are small, so that the increments in lift and moment may bewritten

ACL = CL"lia + CLqq + CLaeASeliCm = Cm Aa + Cm q + Cm_ lifJea q ve

where, in order to maintain a nondimensional form of equations, we have introducedthe dimensionless pitch rate q = qc/2V, and CLq :::::dCJdq, Cmq :::::dCmldq. The q de-

(3.1,2)

(3.1,3)

IThe load factor is the ratio of lift to weight, n = LIW. It is unity in straight horizontal flight.

60

Page 71: Stability Control EtkinReid

3.1 Maneullerability-Elellator Angle per g 61

o

R

(I" = In-1lg

v

nW

W

Figure 3.1 Airplane in a pull-up.

rivatives are discussed in Sec. 5.4. In this form, these equations apply to any configu-ration. From (3.1,1) we get

gcqA= (n - 1)-

2V2

which is more conveniently expressed in terms of the weight coefficient Cw and themass ratio JL (see Sec. 3.15), that is,

Cwq = (n - 1)-2JL

Since the curved flight condition is also assumed to be steady, that is, without angularacceleration, then IlCm = O.Finally, we can relate IlCL to n thus:

(3.1,4)

nW-WIlCL = \pV2S = (n - I)Cw

Equations (3.1,2 and 3) therefore become

(3.1,5)

Page 72: Stability Control EtkinReid

62 Chapter 3. Static Stability and Control-Part 2

which are readily solved for ~a and ~5e to yield the elevator angle per g

~5e Cw [1 ]--1 = - -d, c; - -2 (Clem - CL Cm)n - et a /L -q a a q

~a 1 ( Cw ss, )---- C -C --C --n - 1 - CLa W Lq 2/L La. n - 1

where det is the same expression previously given in (2.4,13d). As has been shown inSec. 2.4 det does not depend on CG position, hence the variation of ~5j(n - 1) withh is provided by the terms in the numerator. Writing Cma = CLJh - hn) (3.1,6a) be-comes

(a)(3.1,6)

and (b)

_~_5_e_= _ CwCLa(2/L - CL) (h _ h + __Cmq.....z....-_)n - 1 2p,det n 2/L - CLq

The derivatives CLq and Cmq both in general vary with h, the former linearly, the latterquadratically, (see Sec. 5.4). Thus (3.1,7), although it appears to be linear in h, is notexactly so. For airplanes with tails, CLq can usually be neglected altogether whencompared with 2/L, and the variation of Cm with h is slight. The equation is then very

q

nearly linear with h, as illustrated in Fig. 3.2. For tailless airplanes, the variation mayshow more curvature. The point where ~5j(n - 1) is zero is called the control:fixedmaneuver point, and is denoted by hm, as shown. From (3.1,7) we see that

(3.1,7)

(3.1,8)

The difference (hm - h) is known as the control-fixed maneuver margin.

o CG position. h

Control-fixed neutral point

lOWe."Control-fixedmaneuver point

Figure 3.2 Elevator angle per g.

Page 73: Stability Control EtkinReid

3.2 Control Force per g 63

3.2 Control Forceper g

From (2.8,4) we get the incremental control force

M> = GSi:eipy2 sc; (3.2,1)

Che is given for rectilinear flight by (2.5,2). Since it too will in general be influencedby q, we write for the incremental value (dB, = 0)

dChe = Che"da + Chell + b2dBe (3.2,2)

The derivative Che is discussed in Sec. 5.4. Using (3.1,4) and (3.1,6b), (3.2,2) isq

readily expanded to give

sc C dB ( C C )__ h_e = __w_ [(2 _ C)C + C C ] + __ e_ b _ L8e he" (3.2 3)1 2 C JL Lq he" hq Lac 1 2 C '

n- JLL" n- L"

From (2.6,4b) we note that the last parenthetical factor is b2C~/CL" or b2a'la. FordBe we use the approximation (3.1,9) in the interest of simplicity and the result fordChe after some algebraic reduction is

sc; Cw a'b2-- = - - - (2JL - CL )(h - h:")n - 1 2JL det q

(3.2,4)

h' = h + det (Che" + Cheq )m m a'b2 CL" 2JL - CLq

In keeping with earlier nomenclature, h:" is the control-free maneuver point and(h:" - h) is the corresponding margin. On noting that Cwipy2 is the wing loading w,we find the control force per g is given by

where (3.2,5)

M> a'bQ = -- = -GS c w __ 2 (2/1 - CL )(h - h' )n - 1 e e 2p..det r- q m

Note that this result applies to both tailed and tailless aircraft provided that the appro-priate derivatives are used. The following conclusions may be drawn from (3.2,6).

(3.2,6)

1. The control force per g increases linearly from zero as the CG is moved for-ward from the control-free maneuver point, and reverses sign for h > h:".

2. It is directly proportional to the wing loading. High wing loading produces"heavier" controls.

3. For similar aircraft of different size but equal wing loading, Q DC Sece; i.e. tothe cube of the linear size.

4. Neither CL nor Yenters the expression for Q explicitly. Thus, apart from Mand Reynolds number effects, Q is independent of speed.

5. The factor JLwhich appears in (3.2,5) causes the separation of the control-freeneutral and maneuver points to vary with altitude, size, and wing loading, inthe same manner as the interval (hm - hn).

Figure 3.3 shows a typical variation of Q with CG position. The statement madeabove that the control force per g is "reversed" when h > h:" must be interpreted cor-

Page 74: Stability Control EtkinReid

64 Chapter 3. Static Stability and Control-Part 2

Control-freemaneuver point-point of zero controlforce per g

Control force per g.

--::+-------:::...=~~;:::------+1&o CG position......•.Control-free neutral point-

point of zero gradient ofcontrol force at

hands-off speed

Figure 3.3

rectly. In the first place this does not necessarily mean a reversal of control move-ment per g, for this is governed by the elevator angle per g. If h;" < h < hm, thenthere would be reversal of Q without reversal of control movement. In the secondplace, the analysis given applies only to the steady state at load factor n, and throwsno light whatsoever on the transition between unaccelerated flight and the pull-upcondition. No matter what the value of h, the initial control force and movement re-quired to start the maneuver will be in the normal direction (backward for a pull-up),although one or both of them may have to be reversed before the final steady state isreached.

CONTROL-FORCE GADGETS

The control forces on a manually controlled airplane can be made to deviate from the"natural" pattern that flows from the size of the airplane, the aerodynamic design,and the speed and altitude of flight without necessarily using either powered controlsor aerodynamic tabs (see Sec. 2.7). Some "gadgets" that can be used to this end aresprings, weights, and variable-ratio sprockets and linkages. These can have the effectof modifying the control-force to trim and the control-force per g, giving the pilot thesame feel as if the control-free neutral and maneuver points were moved. Some de-tails of these effects are given in Sec. 7.1 of Etkin (1972).

3.3 Influence of High-Lift Devices on Trim andPitch Stiffness

Conventional airplanes utilize a wide range of aerodynamic devices for increasingCLmax' These include various forms of trailing edge elements (plain flaps, split flaps,

Page 75: Stability Control EtkinReid

3.3 Influence of High-Lift Devices on Trim and Pitch Stiffness 6S

slotted flaps, etc.), leading edge elements (drooped nose, slats, slots, etc.) and purelyfluid mechanical solutions such as boundary layer control by blowing. Each of thesehas its own characteristic effects on the lift and pitching moment curves, and it is notfeasible to go into them in depth here. The specific changes that result from the "con-figuration-type" devices, i.e. flaps, slots, etc., can always be incorporated by makingthe appropriate changes to hnWb' Cmacwb and CLWb in (2.2,4) and following through theconsequences. Consider for example the common case of part-span trailing edgeflaps on a conventional tailed airplane. The main aerodynamic effects of such flapsare illustrated in Fig. 3.4.2

1. Their deflection distorts the shape of the spanwise distribution of lift on thewing, increasing the vorticity behind the flap tips, as in (a).

2. They have the same effect locally as an increase in the wing-section camber,that is, a negative increment in Cmac and a positive increment in CLwb•

3. The downwash at the tail is increased; both Eo and dE/da will in generalchange.

The change in wing-body Cm is obtained from (2.2,4) as

!i.Cmwb = !i.Cmacwb + !i.CLwb(h - hnwJ

The change in airplane CL is

(3.3,1)

(3.3,2)

and the change in tail pitching moment is

!i.Cmt = at VH!i.E (3.3,3)

When the increments !i.Cmacwb and !i.CLwb are constant with a and !i.hnWb is negligible,then the only effect on CLa and Cma is that of dE/da, and from (2.3,18) and (2.3,21a)these are

s, dE!i.a = !i.CL = -a - !i. -a t S da (3.3,4)

(3.3,5)

The net result on the CL and Cm curves is obviously very much configuration depen-dent. If the Cm - a relation were as in Fig. 3.4e, then the trim change would be verylarge, from al at 8f = 0 to a2 after flap deflection. The CL at a2 is much larger thanat al and hence if the flap operation is to take place without change of trim speed, adown-elevator deflection would be needed to reduce atrim to a3 (Fig. 3.4c). Thiswould result in a nose-down rotation of the aircraft.

2Note that a is still the angle of attack of the zero-lift line of the basic configuration, and that the liftwith flap deflected is not zero at zero a.

Page 76: Stability Control EtkinReid

66 Chapter 3. Static Stability and Control-Part 2

Vorticity In wakebehind flap tips

(a)

Mean aerodynamic center

4e Tallf c::>-

V (relative wind)

(6)

-.---------- .•.CllCc) Cd)

em

/----.....,......;.....,,.....-----Cll

Ce}Figure 3.4 Effect of part-span flaps. (a) Change of lift distribution and vorticity. (b) Changes inforces and moments. (c) Change in CL- (d) Change in downwash. (e) Change in Cm•

3.4 Influence of the Propulsive System on Trim andPitch Stiffness

The influences of the propulsive system upon trim and stability may be both impor-tant and complex. The range of conditions to be considered in this connection is ex-tremely wide. There are several types of propulsive units in common use-recipro-

Page 77: Stability Control EtkinReid

3.4 Influence of the Propulsive System on Trim and Pitch Stiffness 67

eating-engine-driven propellers, turbojets, turboprops, and rockets, and the variationsin engine-plus-vehicle geometry are very great. The analyst may have to deal withsuch widely divergent cases as a high-aspect-ratio straight-winged airplane with sixwing-mounted counterrotating propellers or a low-aspect-ratio delta with buried jetengines. Owing to its complexity, a comprehensive treatment of propulsive systeminfluences on stability is not feasible. There does not exist sufficient theoretical orempirical information to enable reliable predictions to be made under all the above-mentioned conditions. However, certain of the major effects of propellers and propul-sive jets are sufficiently well understood to make it worth while to discuss them, andthis is done in the following.

In a purely formal sense, of course, it is only necessary to add the appropriate di-rect effects, Cmo and aCm laa in (2.3,21 and 22), together with the indirect effects on

p p

the various wing-body and tail coefficients in order to calculate all the results withpower on.

When calculating the trim curves (i.e., elevator angle, tab angle, and controlforce to trim) the thrust must be that required to maintain equilibrium at the conditionof speed and angle of climb being investigated (see Sec. 2.4). For example (see Fig.2.1), assuming that aT ~ 1

CT = CD + Cwsin 'YCwcos 'Y= CL + CTaT

(a)(3.4,1)

(b)

Solving for Cn we get

CD + CL tan 'YCT=------

1 - aT tan 'Y

Except for very steep climb angles, aT tan 'Y~ 1, and we may write approximately,

(3.4,2)

CT = CD + CL tan 'Y (3.4,3)

Let the thrust line be offset by a distance Zp from the CG (as in Fig. 3.5) and neglect-ing for the moment all other thrust contributions to the pitching moment except Tzp,

we have

z= (CD + CL tan 'Y) ~

c(3.4,4)

~ (relati"ewind)

t~ "1Airplane CGPropeller disk II-l

-f- zp

--t(";hrust .; -----r--------T

Figure 3.5 Forces on a propeller.

Page 78: Stability Control EtkinReid

68 Chapter 3. Static Stability and Controt-Part 2

Now let CD be given by the parabolic polar (2.1,2), so that

(3.4,5)

Strictly speaking, the values of CD and CL in (3.4,4 and 5) are those for trimmedflight, i.e. with {)e = {)etrim' For the purposes of this discussion of propulsion effectswe shall neglect the effects of {)e on CDand Cv and assume that the values in (3.4,5)are those corresponding to {)e = O.The addition of this propulsive effect to the Cmcurve for rectilinear gliding flight in the absence of aeroelastic and compressibilityeffects might then appear as in Fig. 3.6a. We note that the gradient -dC.,/dCL for anyvalue of y > 0 is less than for unpowered flight. If dC.,/dCL is used uncritically as acriterion for stability an entirely erroneous conclusion may be drawn from suchcurves.

1. Within the assumptions made above, the thrust moment Tz; is independent ofa, hence dCm Ida = 0 and there is no change in the NP from that for unpow-

p

ered flight.2. A true analysis of stability when both speed and a are changing requires that

the propulsive system controls (e.g., the throttle) be keptjixed, whereas eachpoint on the curves of Fig. 3.6a corresponds to a different throttle setting. Thisparallels exactly the argument of Sec. 2.4 concerning the elevator trim slope.

CII,la)

o

o

(b)

Figure 3.6 Effect of direct thrust moment on Cm(a)curves (see 2.4,ld). (a) Constant 'Y.(b)Constant thrust and power.

Page 79: Stability Control EtkinReid

3.4 Influence of the Propulsive System on Trim and Pitch Stiffness 69

For in fact, under the stated conditions, the Cm - CL curve is transformed intoa curve of c5etrim vs. V by using the relations c5etrim = -Cm(u)/Cma" and CL =w/ipv2s. The slopes of Cm vs. CL and c5etrim vs. V will vanish together.

If a graph of Cm vs. CL be prepared for fixed throttle, then 'Ywill be a variablealong it, and its gradient dCmldCL is an index of stability, as shown in Sec. 6.4. Thetwo idealized cases of constant thrust and constant power are of interest. If the thrustat fixed throttle does not change with speed, then we easily find

T zpC =-C- (a)mp W L C

and (3.4,6)

ac; Tzp(b)--p =-

dCL weIf the power P is invariant, instead of the thrust, then T = PIV and we find

C =!- J p zp C312mp W 2w C L

dCmp = ~ !-J p zp C112

dCL 2 W 2w C

(c)

(3.4,6)

whence (d)

Thus in the constant thrust case, the power-off Cm - CL graph simply has its slopechanged by the addition of thrust, and in the constant power case the shape ischanged as well. The form of these changes is illustrated in Fig. 3.6b and it is evidentby comparison with 3.6a that the behavior of dCmldCL is quite different in these twosituations.

THE INFLUENCE OF RUNNING PROPELLERS

The forces on a single propeller are illustrated in Fig. 3.5, where up is the angle of at-tack of the local flow at the propeller. It is most convenient to resolve the resultantinto the two components T along the axis, and Np in the plane of the propeller. Themoment associated with T has already been treated above, and does not affect Cma'

That due to N; is

x SAC = C 2. J!...m Np C S (3.4,7)

where CNp = N/ipV'-Sp and Spis the propeller disk area. To get the total ACm for sev-eral propellers, increments such as (3.4,7) must be calculated for each and summed.Theory shows (Ribner, 1945) that for small angles CNp is proportional to up' HenceNp contributes to both Cmopand acm/au. The latter is

aCm s, xp aCN aup--p = - - --p -- (3.4,8)au scaup au

If the propeller were situated far from the flow field of the wing, then au/au wouldbe unity. However, for the common case of wing-mounted tractor propellers with the

Page 80: Stability Control EtkinReid

70 Chapter 3. Static Stability and Control-Part 2

propeller plane close to the wing, there is a strong upwash €p at the propeller. Thus

and

ap = a + €p + constoap O€p-=1+-oa oa

(a)(3.4,9)

(b)

where the constant in (3.4,9a) is the angle of attack of the propeller axis relative tothe airplane zero-lift line. Finally,

oCmp = Sp xp (1 + O€p) oCNp

oa S C oa oap(3.4,10)

Increase of Wing Lift

When a propeller is located ahead of a wing, the high-velocity slipstream causesa distortion of the lift distribution, and an increase in the total lift. This is a principalmechanism in obtaining high lift on so-called deflected slipstream STOL airplanes.For accurate results that allow for the details of wing and flap geometry powered-model testing is needed. However, for some cases there are available theoretical re-sults (Ellis, 1971; Kuhn, 1959; Priestly, 1953) suitable for estimates. Both theory andexperiment show that the lift increment tends to be linear in a for constant Cn andhence has the effect of increasing awb' the lift-curve slope for the wing-body combi-nation. From (2.3,23) this is seen to reduce the effect of the tail on the NP location,and can result in a decrease of pitch stiffness.

Effects on the TailThe propeller slipstream can affect the tail principally in two ways. (1) Depend-

ing on how much if any of the tail lies in it, the effective values of at and a; will expe-rience some increase. (2) The downwash values €o and o€/oa may be appreciably al-tered in any case. Methods of estimating these effects are at best uncertain, andpowered-model testing is needed to get results with engineering precision for mostnew configurations. However, some empirical methods (Smelt and Davies, 1937;Millikan, 1940; Weil and Sleeman, 1949) are available that are suitable for somecases.

THE INFLUENCE OF JET ENGINES

The direct thrust moment of jet engines is treated as shown at the beginning of thissection, the constant-thrust idealization given in (3.4,6) often being adequate. In addi-tion, however, there is a normal force on jet engines as well as on propellers.

Jet Normal ForceThe air that passes through a propulsive duct experiences, in general, changes in

both the direction and magnitude of its velocity. The change in magnitude is the prin-cipal source of the thrust, and the direction change entails a force normal to the thrust

Page 81: Stability Control EtkinReid

3.4 Influence of the Propulsive System on Trim and Pitch Stiffness 71

line. The magnitude and line of action of this force can be found from momentumconsiderations. Let the mass flow through the duct be m' and the velocity vectors atthe inlet and outlet be Vi and Vj' Application of the momentum principle then showsthat the reaction on the airplane of the air flowing through the duct is

F = -m'(Vj - V) + F'

where F' is the resultant of the pressure forces acting across the inlet and outlet areas.For the present purpose, F' may be neglected, since it is approximately in the direc-tion of the thrust T. The component of F normal to the thrust line is then found as inFig. 3.7. It acts through the intersection of Vi and Vj' The magnitude is given by

N, = m'Visin 0

or, for small angles,

~=m'ViO (3.4,11)

In order to use this relation, both Vi and 0 are required. It is assumed that Vi has thatdirection which the flow would take in the absence of the engine; that is, 0 equals theangle of attack of the thrust line (Xj plus the upwash angle due to wing induction €j"

(3.4,12)

It is further assumed that the magnitude Vi is determined by the mass flow and inletarea; thus

m'v=-I Api

(3.4,13)

where Ai is the inlet area, and Pi the air density in the inlet. We then get for N, the ex-pression

The corresponding pitching-moment coefficient is

(3.4,14)

Figure 3.7 Momentum change of engine air.

Page 82: Stability Control EtkinReid

72 Chapter 3. Static Stability and Control-Part 2

Figure 3.8 Jet-induced inflow.

Since the pitching moment given by (3.4,14) varies with a at constant thrust, thenthere is a change in Cma given by

sc.; = ~: ip~SC [Xi(1 + ~~) + ()~~] (3.4,15)

The quantities m' and Pi can be determined from the engine performance data, andfor subsonic flow, dE/da is the same as the value dE/da used for propellers. dX/dacan be calculated from the geometry.

Jet Induced InflowA spreading jet entrains the air that surrounds it, as illustrated in Fig. 3.8, thereby

inducing a flow toward the jet axis. If a tailplane is placed in the induced flow field,the angle of attack will be modified by this inflow. A theory of this phenomenonwhich allows for the curvature of the jet due to angle of attack has been formulatedby Ribner (1946). This inflow at the tail may vary with a sufficiently to reduce thestability by a significant amount.

3.5 Effect of Structural FlexibilityMany vehicles when flying near their maximum speed are subject to important aero-elastic phenomena. Broadly speaking, we may define these as the feedback effectsupon the aerodynamic forces of changes in the shape of the airframe caused by theaerodynamic forces. No real structure is ideally rigid, and aircraft are no exception.Indeed the structures of flight vehicles are very flexible when compared with bridges,buildings, and earthbound machines. This flexibility is an inevitable characteristic ofstructures designed to be as light as possible. The aeroelastic phenomena which resultmay be subdivided under the headings static and dynamic. The static cases are thosein which we have steady-state distortions associated with steady loads. Examples areaileron reversal, wing divergence, and the reduction of longitudinal stability. Dy-namic cases include buffeting and flutter. In these the time dependence is an essentialelement. From the practical design point of view, the elastic behavior of the airplaneaffects all three of its basic characteristics: namely performance, stability, and struc-

Page 83: Stability Control EtkinReid

3.5 Effect of Structural Flexibility 73

tural integrity. This subject occupies a well-established position as a separate branchof aeronautical engineering. For further information the reader is referred to one ofthe books devoted to it (Bisplinghoff, 1962; Dowell, 1994).

In this section we take up by way of example a relatively simple aeroelastic ef-fect; namely, the influence of fuselage flexibility on longitudinal stiffness and con-trol. Assume that the tail load L, bends the fuselage so that the tail rotates through theangle ~at = -kLt (Fig. 3.9) while the wing angle of attack remains unaltered. Thenet angle of attack of the tail will then be

at = awb - E - it - kLt

and the tail lift coefficient at 5e = 0 will be

CL, = atat = a,(awb - E - it - kLt)

But L, = CL,!p"Vlst, from which

CL, = a,(awb - E - it - kCL,ip"VlSt) (3.5,1)

Solving for CLt' we get

at .CL, = ----p- (awb - E - It)

1+ katSt"2V2

Comparison of (3.5,2) with (2.3,13) shows that the tail effectiveness has been re-duced by the factor 1/[1 + ka,(p/2)V2St]. The main variable in this expression is V,and it is seen that the reduction is greatest at high speeds. From (2.3,23) we find thatthe reduction in tail effectiveness causes the neutral point to move forward. The shiftis given by

(3.5,2)

~at - ( OE)~h=-V 1--n a H oa (3.5,3)

where

~at = at ( 1 + ka\PV2St - 1) (3.5,4)

The elevator effectiveness is also reduced by the bending of the fuselage. For, ifwe consider the case when 5e is different from zero, then (3.5,1) becomes

CL, = a,(awb - E - it - kCL,!pV2St) + ae5e

Figure 3.9 Tail rotation due to fuselage bending.

Page 84: Stability Control EtkinReid

74 Chapter 3. Static Stability and Control-Part 2

and (3.5,2) becomes

a,(Cl!wb - e - it) + ae5eCL, = --------1 + kat~py2St

Thus the same factor 1/(1 + katp/2lftSt) that operates on the tail lift slope at also mul-tiplies the elevator effectiveness ae•

3.6 GroundEffectAt landing and takeoff airplanes fly for very brief (but none the less extremely impor-tant) time intervals close to the ground. The presence of the ground modifies the flowpast the airplane significantly, so that large changes can take place in the trim and sta-bility. For conventional airplanes, the takeoff and landing cases provide some of thegoverning design criteria.

The presence of the ground imposes a boundary condition that inhibits the down-ward flow of air normally associated with the lifting action of the wing and tail. Thereduced downwash has three main effects, being in the usual order of importance:

1. A reduction in s, the downwash angle at the tail.2. An increase in the wing-body lift slope awb'

3. An increase in the tail lift slope at.

The problem of calculating the stability and control near the ground then resolves it-self into estimating these three effects. When appropriate values of i)e1i)a, awb' and athave been found, their use in the equations of the foregoing sections will readilyyield the required information. The most important items to be determined are the el-evator angle and control force required to maintain CLmsx in level flight close to theground. It will usually be found that the ratio a.Ia is decreased by the presence of theground. (2.3,23) shows that this would tend to move the neutral point forward. How-ever, the reduction in i)e1i)a is usually so great that the net effect is a large rearwardshift of the neutral point. Since the value of Cmo is only slightly affected, it turns outthat the elevator angle required to trim at C£max is much larger than in flight remotefrom the ground. It commonly happens that this is the critical design condition on theelevator, and it will govern the ratio S,jSt, or the forward CG limit (see Sec. 3.7).

3.7 CGLimitsOne of the dominant parameters of longitudinal stability and control has been shownin the foregoing sections to be the fore-and-aft location of the CG (see Figs. 2.14,2.18,2.19,2.25,2.27,2.28, and 3.2). The question now arises as to what range of CGposition is consistent with satisfactory handling qualities. This is a critical designproblem, and one of the most important aims of stability and control analysis is toprovide the answer to it. Since aircraft always carry some disposable load (e.g., fuel,armaments), and since they are not always loaded identically to begin with (varia-tions in passenger and cargo load), it is always necessary to cater for a variation inthe CG position. The range to be provided for is kept to a minimum by proper loca-

Page 85: Stability Control EtkinReid

3.7 CGlimits 75

tion of the items of variable load, but still it often becomes a difficult matter to keepthe handling qualities acceptable over the whole CG range. Sometimes the problem isnot solved, and the airplane must be subjected to restrictions on the fore-and-aft dis-tribution of its variable load when operating at part load.

THE AFT LIMIT

The most rearward allowable location of the CO is determined by considerations oflongitudinal stability and control sensitivity. The behavior of the five principal con-trol gradients are summarized in Fig. 3.10 for the case when the aerodynamic coeffi-cients are independent of speed. From the handling qualities point of view, none ofthe gradients should be "reversed," that is, they should have the signs associated withlow values of h. When the controls are reversible, this requires that h < h~. If thecontrols are irreversible, and if the artificial feel system is suitably designed, then thecontrol force gradient dPldV can be kept negative to values of h > h~, and the rearlimit can be somewhat farther back than with reversible controls. The magnitudes ofthe gradients are also important. If they are allowed to fall to very small values thevehicle will be too sensitive to the controls. When the coefficients do not depend onspeed, as assumed for Fig. 3.10, the NP also gives the stability boundary (this isproved in Chap. 6), the vehicle becoming unstable for h > h~with free controls orh > hn with fixed controls. If the coefficients dependent on speed, for example, Cm =Cm(M), then the CO boundary for stability will be different and may be forward oftheNP.

":;;o:+-----+:::0..-::'--.h ~O:+---.--;~~--·h

A6.ii"=1

Q

o

Figure 3.10 The five control gradients.

Page 86: Stability Control EtkinReid

76 Chapter 3. Static Stability and Control-Part 2

As noted in Chap. 1, it is possible to increase the inherent stability of a flight ve-hicle. Stability augmentation systems (SAS) are in widespread use on a variety of air-planes and rotorcraft. If such a system is added to the longitudinal controls of an air-plane, it permits the use of more rearward CG positions than otherwise, but the riskof failure must be reckoned with, for then the airplane is reduced to its "inherent" sta-bility, and would still need to be manageable by a human pilot.

THE FORWARD LIMIT

As the CG moves forward, the stability of the airplane increases, and larger controlmovements and forces are required to maneuver or change the trim. The forward CGlimit is therefore based on control considerations and may be determined by anyoneof the following requirements:

1. The control force per g shall not exceed a specified value.2. The control-force gradient at trim, aPIa V, shall not exceed a specified value.3. The control force required to land, from trim at the approach speed, shall not

exceed a specified value.4. The elevator angle required to land shall not exceed maximum up elevator.5. The elevator angle required to raise the nose-wheel off the ground at takeoff

speed shall not exceed the maximum up elevator.

3.8 Lateral AerodynamicsIn the preceding sections of this and the previous chapter we discussed aerodynamiccharacteristics of symmetrical configurations flying with the velocity vector in theplane of symmetry. As a result the only nonzero motion variables were V, a, and q,and the only nonzero forces and moments were T, D, L, and M. We now tum to thecases in which the velocity vector is not in the plane of symmetry, and in which yaw-ing and rolling displacements (f3, 4» are present. The associated force and momentcoefficients are C; C/, and Cn•

One of the simplifying aspects of the longitudinal motion is that the rotation isabout one axis only (the y axis), and hence the rotational stiffness about that axis is avery important criterion for the dynamic behavior. This simplicity is lost when we goto the lateral motions, for then the rotation takes place about two axes (x and z). Themoments associated with these rotations are cross-coupled, that is, roll rotation p pro-duces a yawing moment C; as well as rolling moment C/, and yaw displacement f3and rate r both produce rolling and yawing moments. Furthermore, the roll and yawcontrols are also often cross-coupled--deflection of the ailerons can produce signifi-cant yawing moments, and deflection of the rudder can produce significant rollingmoments.

Another important difference between the two cases is that in "normal" flight-that is, steady rectilinear symmetric motion, all the lateral motion and force variablesare zero. Hence there is no fundamental trimming problem-the ailerons and rudderwould be nominally undeflected. In actuality of course, these controls do have a sec-ondary trimming function whenever the vehicle has either geometric or inertial asym-metries-for example, one engine off, or multiple propellers all rotating the same

Page 87: Stability Control EtkinReid

3.9 Weathercock Stability (Yaw Stiffness) 77

way. Because the gravity vector in normal flight also lies in the plane of symmetry,the CG position is not a dominant parameter for the lateral characteristics as it is forthe longitudinal. Thus the CG limits, (see Sec. 3.7) are governed by considerationsderiving from the longitudinal characteristics.

3.9 Weathercock Stability (Yaw Stiffness)Application of the static stability principle to rotation about the z axis suggests that astable airplane should have "weathercock" stability. That is, when the airplane is atan angle of sideslip j3 relative to its flight path (see Fig. 3.11), the yawing momentproduced should be such as to tend to restore it to symmetric flight. The yawing mo-ment N is positive as shown. Hence the requirement for yaw stiffness is that aNlaj3must be positive. The nondimensional coefficient of N is

Nc; = ipV2Sb

and hence for positive yaw stiffness aC,liJj3 must be positive. The usual notation forthis derivative is

(3.9,1)

This quantity is analogous in some respects to the longitudinal stability parameterCm". It is estimated in a similar way by synthesis of the contributions of the variouscomponents of the airplane. The principal contributions are those of the body and thevertical-tail surface. By contrast with Cm", the wing has little influence in most cases,and the CG location is a weak parameter. Whether or not a positive value of Cn/l willproduce lateral stability can only be determined by a full dynamic analysis such as isdone in Chap. 6.

v

\,Figure 3.11 Sideslip angle and yawing moment.

Page 88: Stability Control EtkinReid

78 Chapter 3. Static Stability and Control-Part 2

In Fig. 3.12 are shown the relevant geometry and the lift force LF acting on thevertical tail surface. If the surface were alone in an airstream, the velocity vector VF

would be that of the free stream, so that (cf. Fig. 3.11) (XF would be equal to -/3.When installed on an airplane, however, changes in both magnitude and direction ofthe local flow at the tail take place. These changes may be caused by the propellerslipstream, and by the wing and fuselage when the airplane is yawed. The angular de-flection is allowed for by introducing the sidewash angle (T, analogous to the down-wash angle E. a is positive when it corresponds to a flow in the y direction; that is,when it tends to increase (XF' Thus the angle of attack is

(3.9,2)

and the lift coefficient of the vertical-tail surface is

(3.9,3)

The lift is then

(3.9,4)

and the yawing moment is

-----~~-----y-----yCG

IF

Mean aerodynamiccenter of fin andrudder

LF

~H,

Figure 3.12 Vertical-tail sign conventions.

Page 89: Stability Control EtkinReid

3.9 Weathercock Stability (Yaw Stiffness) 79

SFIF (VF)2thus CnF = -CLF ----.sb"V (3.9,5)

The ratio S~~Sb is analogous to the horizontal-tail volume ratio, and is thereforecalled the vertical-tail volume ratio, denoted here by Vy. Equation 3.9,5 then reads

CnF = -VyCLF (~rand the corresponding contribution to the weathercock stability is

d;F = -Vy(~r d~;F= VyaF(~r(l- ~;)THE SIDEWASH FACTOR iJuliJ{J

(3.9,6)

(3.9,7)

Generally speaking, the sidewash is difficult to estimate with engineering precision.Suitable wind-tunnel tests are required for this purpose. The contribution from thefuselage arises through its behavior as a lifting body when yawed. Associated withthe side force that develops is a vortex wake which induces a lateral-flow field at thetail. The sidewash from the propeller is associated with the side force which actsupon it when yawed, and may be estimated by the method of (Ribner, 1944). Thecontribution from the wing is associated with the asymmetric structure of the flowwhich develops when the airplane is yawed. This phenomenon is especially pro-nounced with low-aspect-ratio swept wings. It is illustrated in Fig. 3.13.

Netinducedflow

Figure 3.13 Vortex wake of yawed wing.

Page 90: Stability Control EtkinReid

80 Chapter 3. StatU:Stability and Control-Part 2

THE VELOCITY RATIO VFIV

When the vertical tail is not in a propellor slipstream, VFIV is unity. When it is in aslipstream, the effective velocity increment may be dealt with as for a horizontal tail.

CONTRIBUTION OF PROPELLER NORMAL FORCE

The yawing moment produced by the normal force that acts on the yawed propeller iscalculated in the same way as the pitching-moment increment dealt with in Sec. 3.4.The result is similar to (3.4,8):

d aCn = _ xp Sp acNpaa a . O~~J-' b S cxp

This is known as the propeller fin effect and is negative (i.e., destabilizing) when thepropeller is forward of the eG, but is usually positive for pusher propellers. There isa similar yawing moment effect for jet engines (see Exercise 3.7).

3.10 YawControlIn most flight conditions it is desired to maintain the sideslip angle at zero. If the air-plane has positive yaw stiffness, and is truly symmetrical, then it will tend to fly inthis condition. However, yawing moments may act upon the airplane as a result ofunsymmetrical thrust (e.g., one engine inoperative), slipstream rotation, or the un-symmetrical flow field associated with turning flight. Under these circumstances, f3can be kept zero only by the application of a control moment. The control that pro-vides this is the rudder. Another condition requiring the use of the rudder is thesteady sideslip, a maneuver sometimes used, particularly with light aircraft, to in-crease the drag and hence the glide path angle. A major point of difference betweenthe rudder and the elevator is that for the former trimming the airplane is a secondaryand not a primary function. Apart from this difference, the treatment of the two con-trols is similar. From (3.9,3) and (3.9,6), the rate of change of yawing moment withrudder deflection is given by

ec, (VF)2 ec., (VF)2Cn8r = a8

r= -Vv V a8

r= -arVV V

This derivative is sometimes called the "rudder power." It must be large enough tomake it possible to maintain zero sideslip under the most extreme conditions ofasymmetric thrust and turning flight.

A second useful index of the rudder control is the steady sideslip angle that canbe maintained by a given rudder angle. The total yawing moment during steadysideslip may be written

(3.10,1)

c, = Cnpf3 + Cn8r 8r

For steady motion, en = 0, and hence the desired ratio is

f3 c.;-=---s, c;

(3.10,2)

(3.10,3)

Page 91: Stability Control EtkinReid

3.11 Roll Stiffness 81

The rudder hinge moment and control force are also treated in a manner similarto that employed for the elevator. Let the rudder hinge-moment coefficient be givenby

Chr = b, aF + b2Sr

The rudder pedal force will then be given by

P = G ~ V~SrCr(bI aF + b28r)

(3.10,4)

(3.10,5)

where G is the rudder system gearing.The effect of a free rudder on the directional stability is found by setting Chr = °

in (3.10,4). Then the rudder floating angle is

(3.10,6)

The vertical-tail lift coefficient with rudder free is found from (3.9,3) to be

blC~F = aFaF - a; b

2aF

(a; bl)=aFaF 1---aF bz

The free control factor for the rudder is thus seen to be of the same form as that forthe elevator (see Sec. 2.6) and to have a similar effect.

(3.10,7)

3.11 Roll StiffnessConsider a vehicle constrained, as on bearings in a wind tunnel, to one degree offreedom-rolling about the x axis. The forces and moments resulting from a fixed dis-placement c/> are fundamentally different in character from those associated with therotations a and {3.In the first place if the x axis coincides with the velocity vector V,no aerodynamic change whatsoever follows from the fixed rotation c/> (see Fig. 3.14).The aerodynamic field remains symmetrical with respect to the plane of symmetry,the resultant aerodynamic force remains in that plane, and no changes occur in any ofthe aerodynamic coefficients. Thus the roll stiffness dC/de/> = Cz</> is zero in that case.

If the x axis does not coincide with V, then a second-order roll stiffness resultsthrough the medium of the derivative aC/d{3 = Cz13• Let the angle of attack of the xaxis be ax (see Fig. 1.7), then the velocity vector when e/>= 0 is

[vcos ax]

VI = 0V sin ax

(3.11,1)

After rolling through angle e/>about Ox, the x component of the velocity vector re-mains unchanged, but the component V sin ax has projections on both of the new y

Page 92: Stability Control EtkinReid

82 Chapter 3. Static Stability and Control-Part 2

Rolling moment, L

z

W(weight)

Figure 3.14 Rolled airplane.

and z axes. Thus there is now a sideslip, and hence, an angle {3and a resulting rollingmoment. Using the notation of Appendix A.4, we get for the velocity vector in thenew reference frame after the rotation c/J

[Vcos ax ]

V 2 = L1(c/J)V1 = V sin ax sin c/JV sin ax cos c/J

(3.11,2)

Thus the sideslip component is v = V sin ax sin c/J, and the sideslip angle is

v{3= sin" V = sin" (sin ax sin c/J)

As a result of this positive {3,and the usually negative ClfJ there is a restoring rollingmoment Clj3, that is,

(3.11,3)

l:1CI = ClfJ sin" (sin ax sin c/J)

For small ax, we get the approximate result

l:1CI == ClfJ sin-1 (ax sin c/J) == ClfJax sin c/J

and if c/J also is small,

(a)

(b) (3.11,4)

l:1CI == ClfJaxc/J

The stiffness derivative for rolling about Ox is then from (3.11,4a)

aCI sin ax cos c/JiJc/J = ClfJ (1 - sirr' ax sin2 c/J)112

(c)

(a)

or for ax ~ 1,

(b) (3.11,5)

Page 93: Stability Control EtkinReid

3.12 The Derivative C.p 83

ec,acfJ == C1,Px

Thus there is a roll stiffness that resists rolling if ax is >0, and would tend to keep thewings level. If rolling occurs about the wind vector, the stiffness is zero and the vehi-cle has no preferred roll angle. If ax < 0, then the stiffness is negative and the vehiclewould roll to the position cfJ = 180°, at which point C1 = 0 and C1q, < O.

The above discussion applies to a vehicle constrained, as stated, to one degree offreedom. It does not, by any means, give the full answer for an unconstrained air-plane to the question: "What happens when the airplane rolls away from a wings-level attitude-does it tend to come back or not?" That answer can only be providedby a full dynamic analysis like the kind given in Chaps. 6 and 7. The roll stiffness ar-gument given above, however, does help in understanding the behavior of slender air-planes, ones with very low aspect ratio and hence small roll inertia. These tend, in re-sponse to aileron deflection when at angle of attack, to rotate about the x axis, not thevelocity vector, and hence experience the roll stiffness effect at the beginning of theresponse.

Even though airplanes have no first-order aerodynamic roll stiffness, stable air-planes do have an inherent tendency to fly with wings level. They do so because ofwhat is known as the dihedral effect. This is a complex pattern involving gravity andthe derivative C/~, which owes its existence largely to the wing dihedral (see Sec.3.12). When rolled to an angle cfJ, there is a weight component mg sin cfJ in the y di-rection (Fig. 3.14). This induces a sideslip velocity to the right, with consequent f3 >0, and a rolling moment C1{Jf3 that tends to bring the wings level. The rolling andyawing motions that result from such an initial condition are, however, strongly cou-pled, so no significant conclusions can be drawn about the behavior except by a dy-namic analysis (see Chap. 6).

(c)

3.12 The Derivative C1p

The derivative C1{J is of paramount importance. We have already noted its relation toroll stiffness and to the tendency of airplanes to fly with wings level. The primarycontribution to Cl~is from the wing-its dihedral angle, aspect ratio, and sweep allbeing important parameters.

The effect of wing dihedral is illustrated in Fig. 3.15. With the coordinate systemshown, the normal velocity component Vn on the right wing panel (R) is, for small di-hedral angle I',

Vn = w cos r + v sin r== w + vr

and that on the other panel is w - vr. The terms ± vr IV = ± f3r represent oppositechanges in the angle of attack of the two panels resulting from sideslip. The "up-wind" panel has its angle of attack and therefore its lift increased, and vice versa. Theresult is a rolling moment approximately linear in both f3 and I', and hence a fixed

Page 94: Stability Control EtkinReid

84 Chapter 3. StatU: StabiUty and Control-Part 2

~II Velocity

~ components/' W of wing

:c

II

Figure 3.15 Dihedral effect. Vn = normal velocity of panel R = w cos r + v sin r '* w + vr :.vr vf3r

!!.a of R due to dihedral '*V =V = f3r.

value of C1fJ for a given r.This part of C1fJ is essentially independent of wing angle ofattack so long as the flow remains attached.

Even in the absence of dihedral, a flat lifting wing panel has a CZfJ proportional toCv Consider the case of Fig. 3.16. The vertical induced velocity (downwash) of thevortex wake is greater at L than at R simply by virtue of the geometry of the wake.Hence the local wing angle of attack and lift are less at L than at R, and a negative C1

results. Since this effect depends, essentially linearly, on the strength of the vortexwake, which is itself proportional to the wing Cv then the result is fi.C1fJ DC Cv

\\'("

\\\

V (relative wind)

Figure 3.16 Yawed lifting wing.

Page 95: Stability Control EtkinReid

3.12 The Derivative ell' 8S

---- ~£H"'''~ ..'!L

-----~~~w~;Figure 3.17 Influence of body on CI".

INFLUENCE OF FUSELAGE ON C1fJ

The flow field of the body interacts with the wing in such a way as to modify its di-hedral effect. To illustrate this, consider a long cylindrical body, of circular cross sec-tion, yawed with respect to the main stream. Consider only the cross-flow componentof the stream, of magnitude V{3, and the flow pattern which it produces about thebody. This is illustrated in Fig. 3.17. It is clearly seen that the body induces verticalvelocities which, when combined with the mainstream velocity, alter the local angleof attack of the wing. When the wing is at the top of the body (high-wing), then theangle of attack distribution is such as to produce a negative rolling moment; that is,the dihedral effect is enhanced. Conversely, when the airplane has a low wing, the di-hedral effect is diminished by the fuselage interference. The magnitude of the effectis dependent upon the fuselage length ahead of the wing, its cross-section shape, andthe planform and location of the wing. Generally, this explains why high-wing air-planes usually have less wing dihedral than low-wing airplanes.

INFLUENCEOFSWEEPONC~

Wing sweep is a major parameter affecting C1fJ• Consider the swept yawed wing illus-trated in Fig. 3.18. According to simple sweep theory it is the velocity Vn normal to awing reference line (~ chord line for subsonic, leading edge for supersonic) that de-termines the lift. It follows obviously that the lift is greater on the right half of the

Figure 3.18 Dihedral effect of a swept wing.

Page 96: Stability Control EtkinReid

86 Chapter 3. Static Stability and Control-Part 2

v:•••~~ .•. -<- __

CG

Figure 3.19 Dihedral effect of the vertical tail.

wing shown than on the left half, and hence that there is a negative rolling moment.The rolling moment would be expected for small {3to be proportional to

CL[(V;)right - (V;)left] = CLV2[COS2 (A - (3) - cos? (A + (3)]== 2CL{3VZ sin 2A

The proportionality with CL and {3is correct, but the sin 2A factor is not a good ap-proximation to the variation with A. The result is a C1{J ee CL' that can be calculatedby the methods of linear wing theory.

INFLUENCE OF FIN ON C1{J

The sideslipping airplane gives rise to a side force on the vertical tail (see Sec. 3.9).When the mean aerodynamic center of the vertical surface is appreciably offset fromthe rolling axis (see Fig. 3.19) then this force may produce a significant rolling mo-ment. We can calculate this moment from (3.9,3). When the rudder angle is zero, thatis, B, = 0, the moment is found to be

thus

and

(3.12,1)

3.13 Ron ControlThe angle of bank of the airplane is controlled by the ailerons. The primary functionof these controls is to produce a rolling moment, although they frequently introduce ayawing moment as well. The effectiveness of the ailerons in producing rolling andyawing moments is described by the two control derivatives ac;aSa and aCiaSa' Theaileron angle Sa is defined as the mean value of the angular displacements of the twoailerons. It is positive when the right aileron movement is downward (see Fig. 3.20).The derivative ac/asa is normally negative, right aileron down producing a roll tothe left.

Page 97: Stability Control EtkinReid

3.13 RoO Control 87

Figure 3.20 Aileron angle.

For simple flap-type ailerons, the increase in lift on the right side and the de-crease on the left side produce a drag differential that gives a positive (nose-right)yawing moment. Since the normal reason for moving the right aileron down is to ini-tiate a tum to the left, then the yawing moment is seen to be in just the wrong direc-tion. It is therefore called aileron adverse yaw. On high-aspect-ratio airplanes thistendency may introduce decided difficulties in lateral control. Means for avoidingthis particular difficulty include the use of spoilers and Frise ailerons. Spoilers are il-lustrated in Fig. 3.21. They achieve the desired result by reducing the lift and increas-ing the drag on the side where the spoiler is raised. Thus the rolling and yawing mo-ments developed are mutually complementary with respect to turning. Frise aileronsdiminish the adverse yaw or eliminate it entirely by increasing the drag on the side ofthe upgoing aileron. This is achieved by the shaping of the aileron nose and thechoice of hinge location. When deflected upward, the gap between the control sur-face and the wing is increased, and the relatively sharp nose protrudes into thestream. Both these geometrical factors produce a drag increase.

--y~- Section through spoiler

Figure 3.21 Spoilers.

Page 98: Stability Control EtkinReid

88 Chapter 3. Static Stability and Control-Part 2

The deflection of the ailerons leads to still additional yawing moments once theairplane starts to roll. These are caused by the altered flow about the wing and tail.These effects are discussed in Sec. 5.7 (Cn), and are illustrated in Figs. 5.12 and5.15.

A final remark about aileron controls is in order. They are functionally distinctfrom the other two controls in that they are rate controls. If the airplane is restrictedonly to rotation about the x axis, then the application of a constant aileron angle re-sults in a steady rate of roll. The elevator and rudder, on the other hand, are displace-ment controls. When the airplane is constrained to the relevant single axis degree offreedom, a constant deflection of these controls produces a constant angular displace-ment of the airplane. It appears that both rate and displacement controls are accept-able to pilots.

AILERON REVERSAL

There is an important aeroelastic effect on roll control by ailerons that is significanton most conventional airplanes at both subsonic and supersonic speeds. This resultsfrom the elastic distortion of the wing structure associated with the aerodynamic loadincrement produced by the control. As illustrated in Fig. 2.23, the incremental loadcaused by deflecting a control flap at subsonic speeds has a centroid somewhere nearthe middle of the wing chord. At supersonic speeds the control load acts mainly onthe deflected surface itself, and hence has its centroid even farther to the rear. If thisload centroid is behind the elastic axis of the wing structure, then a nose-down twistof the main wing surface results. The reduction of angle of attack corresponding to 5a

> 0 causes a reduction in lift on the surface as compared with the rigid case, and aconsequent reduction in the control effectiveness. The form of the variation of CI6awith ~pV2 for example can be seen by considering an idealized model of the phenom-enon. Let the aerodynamic torsional moment resulting from equal deflection of thetwo ailerons be T(y) oc ~pVZ5a and let T(y) be antisymmetric, T(y) = -T( -y). Thetwist distribution corresponding to T(y) is O(y), also antisymmetric, such that O(y) isproportional to T at any reference station, and hence proportional to ~pV25a. Finally,the antisymmetric twist causes an antisymmetric increment in the lift distribution,giving a proportional rolling moment increment tiCI = kipVZ5a. The total rolling mo-ment due to aileron deflection is then

tiCI = (CI8)rigid5a + ~pV25a

and the control effectiveness is

(3.13,1)

Cl8a= (CI8)rigid + ~pVZ (3.13,2)

As noted above, (CI8)rigid is negative, and k is positive if the centroid of the aileron-induced lift is aft of the wing elastic axis, the common case. Hence Icl6a1diminisheswith increasing speed, and vanishes at some speed VR, the aileron reversal speed.Hence

or0= (CI8)rigid + k!pV~k = -(CI8)rigii~p~ (3.13,3)

Page 99: Stability Control EtkinReid

3.14 Exercises 89

Substitution of (3.13,3) into (3.13,2) yields

CIBa = (CIB)rigid (1 - ~~) (3.13,4)

This result, of course, applies strictly only if the basic aerodynamics are not Mach-number dependent, i.e, so long as VR is at a value of M appreciably below 1.0. Other-wise k and (CIB)rigid are both functions of M, and the equation corresponding to(3.13,4) is

k(M) V2

CIBa(M) = (CIB)rigid(M) - k(MR) (CIB)rigiiMR) ~

where MR is the reversal Mach number.It is evident from (3.13,4) that the torsional stiffness of the wing has to be great

enough to keep VR appreciably higher than the maximum flight speed if unacceptableloss of aileron control is to be avoided.

(3.13,5)

3.14 Exercises3.1 Derive (3.1,4).

3.2 Derive an expression for the elevator angle per g in dimensional form. Denote the de-rivatives of L and M with respect to a and q by aUaq = Lq, and so on. There are twochoices: (1) do the derivation in dimensional form from the beginning, or (2) convertthe nondimensional result (3.1,6) to dimensional form. Do it both ways and checkthat they agree.

3.3 Calculate the variation of the control force per g with altitude from the followingdata. Ignore propulsion effects.

Geometric DataWeight, WWing area, SWing mean aerodynamic chord, cItTail area, Sf

SeMean elevator chord, ce

G

Aerodynamic Dataa

50,000 lb (222,500 N)937.5 ft2 (87.10 m2)12.80 ft (3.90 m)31.85 ft (9.71 m)230 ft2 (21.4 m2)71.30 ft2 (6.62 nr')2.21 ft (0.674 m)300/ft (98.4°/m)

0.088/deg0.044/degO.064/dego

Page 100: Stability Control EtkinReid

90 Chapter 3. Static Stability and Control-Part 2

». -0.17/radb2 -OA8/radCh•q -0.846

CLq 0~ -n9(h - hn) -0.10ae\aa 0.30

~ 03.4 A small manually controlled airplane has an undesirable handling characteristic-the

control force per g is too large. List some design changes that could reduce it, and de-scribe the other consequences that each such change would entail.

3.5 The range of elevator motion on an airplane is from 20° down to 30° up. Use Table1.3 as a guide, a fellow student as a model, and a tape measure to arrive at a reason-able value for the elevator gearing ratio G.

3.6 Two airplanes are similar, but one is jet-propelled and the other has a piston engineand propeller. The thrust line in each case is well below the CG with z/c = 0.4. Thepower-off pitching moment at 5e = 0 is Cm = 0.1 - 0.2 CL- The throttle is set to givelevel flight with CL = 004 and UD = 12. Consider several steady rectilinear flightconditions having the same throttle setting but different elevator settings, CL valuesand flight-path angles. Find dCm(a)/dCL (for 5e = 0) for the two airplanes when pass-ing through the altitude corresponding to the level flight conditions. As indicated inSec. 3.4 and (6.4,10) dC,jdCL is an index of static longitudinal stability under certainconditions. Assuming that these conditions are met in this problem, how will the sta-tic longitudinal stability of the two aircraft change as the aircraft slow down?

ac3.7 Derive an expression for the increment A a; attributable to a jet engine. (Hint, re-

fer to (304,15).)

3.8 Suppose that as a result of an accident in flight the rear fuselage of an airplane isdamaged, so that the flexibility parameter k in (3.5,1) et seq. is suddenly increased.The effect is large enough that the pilot notices a loss in longitudinal stability andcontrol. Bearing in mind that the integrity of the fuselage structure depends on thetail load L, and the stability and control on the factor in parentheses in (3.5,4), ana-lyze how the situation changes as the pilot slows down and descends to an emergencylanding. Consider two cases; (1) CL, initially positive, (2) CL, initially negative.

3.9 Derive (3.9,8). Explain clearly each step in the development and justify any assump-tions you make.

3.10 Use Appendix B to determine the elevator hinge moment parameters b, and b2 for aNACA 0009 airfoil (a symmetric airfoil with a thickness-to-chord ratio of tic ::::0.09). The elevator has an elliptic nose, a sealed gap and a balance ratio of 0.2. In us-

ing the curves assume that transition is at the leading edge; R = 107; tan ( ;) =

(4)!m)tan 2 = 0.12; F3 = 1; c,lc = 0.325; A = 4.84; M = O.

Page 101: Stability Control EtkinReid

3.15 Additional Symbols Introduced in Chapter 3 91

3.15 Additional Symbols Introduced in Chapter 3

aF dCddaF

a, dCL/darb span of airplanecr mean rudder chordChr rudder hinge-moment coefficientC, rolling-moment coefficient, UipV2sb

CI/3 dCld/3CLF vertical-tail lift coefficient

C/. dCJdqc.; damping in pitch, dC,jdqC; yawing-moment coefficient, NI!pV2Sb

Cn/3 dCjd/3Cns, dCjo8rc, TI!pV2S

Cw W/ipV2s

d diameter of propeller or jetg acceleration due to gravityhm maneuver point, stick fixed (see 3.1,8)h;" maneuver point, stick free (see 3.2,5)L rolling momentLF vertical-tail lift

IF see Fig. 3.12m airplane mass

Mq dMloq

n load factor, UWN yawing momentq angular velocity in pitch, rad/sq qc/2V

SF vertical-tail areaSp propeller disk areaS, area of rudderT thrust of one propulsion unitVF effective velocity vector at the finVv vertical-tail volume ratio, SplplSb

w wing loading, W/S

ZF see Fig. 3.19

Page 102: Stability Control EtkinReid

92 Chapter 3. Static Stability and Control-Part 2

aF effective angle of attack of the fin, (see Fig. 3.12)f3 sideslip angler dihedral angleEp upwash at propeller5a aileron angle (see Fig. 3.20)B, rudder angle (see Fig. 3.12)A sweepback anglea sidewash angleJ.L 2m/pSc¢ bank angle

Page 103: Stability Control EtkinReid

CHAPTER 4

General Equations ofUnsteady Motion

4.1 General RemarksThe basis for analysis, computation, or simulation of the unsteady motions of a flightvehicle is the mathematical model of the vehicle and its subsystems. An airplane inflight is a very complicated dynamic system. It consists of an aggregate of elasticbodies so connected that both rigid and elastic relative motions can occur. For exam-ple, the propeller or jet-engine rotor rotates, the control surfaces move about theirhinges, and bending and twisting of the various aerodynamic surfaces occur. The ex-ternal forces that act on the airplane are also complicated functions of its shape andits motion. It seems clear that realistic analyses of engineering precision are notlikely to be accomplished with a very simple mathematical model. The model that isdeveloped in the following has been found by aeronautical engineers and researchersto be very useful in practise. We begin by first treating the vehicle as a single rigidbody with six degrees of freedom. This body is free to move in the atmosphere underthe actions of gravity and aerodynamic forces-it is primarily the nature and com-plexity of aerodynamic forces that distinguish flight vehicles from other dynamicsystems. Next we add the gyroscopic effects of spinning rotors and then continuewith a discussion of structural distortion (aeroelastic effects).

As was noted in Chap. 1, the Earth is treated as flat and stationary in inertialspace. These assumptions simplify the model enormously, and are quite acceptablefor dealing with most problems of airplane flight. The effects of a round rotatingEarth are treated at some length in Etkin (1972).

Extensive use is made in the developments that follow of linear algebra, withwhich the reader is assumed to be familiar. Appendix A.l contains a brief review ofsome pertinent material.

4.2 The Rigid-Body EquationsIn the interest of completeness, the rigid-body equations are derived from first princi-ples, that is to say, we apply Newton's laws to an element dm of the airplane, andthen integrate over all elements. The velocities and accelerations must of course berelative to an inertial, or Newtonian, frame of reference. As we noted in Sec. 1.6 theframe FE' fixed to the Earth, is assumed to be such a frame. We also noted there thatvelocities relative to FE are identified by a superscript E. In order to avoid the carry-ing of the cumbersome superscript throughout the following development, we shall

93

Page 104: Stability Control EtkinReid

94 Chapter 4. General Equations of Unsteady Motion

temporarily assume that W = 0 in (1.6,1), so that VE = V, and make an appropriateadjustment at the end. In the frame FB

VB = [u v wf (4.2,1)

The position vector of dm relative to the origin of FE is r, + r (see Fig. 4.1). In theframe FE'

(4.2,2)

and in the frame FB

(4.2,3)

The inertial velocity of dm is

VE= (tcE + tE) = VE + tE

The momentum of dm is vdm, and of the whole airplane is

J vEdm = J (VE + tE)dm = VEJ dm + J tEdm

(4.2,4)

(4.2,5)

ZE

Figure 4.1 Axes.

Page 105: Stability Control EtkinReid

4.2 The Rigid-Body Equations 95

Since C is the mass center, the last integral in (4.2,5) is zero and

(4.2,6)

where m is the total mass of the airplane. Newton's second law applied to dm is

dfE = vEdm (4.2,7)

where dfE is the resultant of all forces acting on dm. The integral of (4.2,7) is

fE = f dfE = f vEdm

or, from (4.2,6)

fE = mVE (4.2,8)

The quantity IdfE is a summation of all the forces that act upon all the elements.The internal forces, that is, those exerted by one element upon another, all occur inequal and opposite pairs, by Newton's third law, and hence contribute nothing to thesummation. Thus fE is the resultant external force acting upon the airplane.

This equation relates the external force on the airplane to the motion of the CG.We need also the relation between the external moment and the rotation of the air-plane. It is obtained from a consideration of the moment of momentum. The momentof momentum of dm with respect to C is by definition db = r X vdm. It is convenientin the following to use the matrix form of the cross product (see Appendix A.l) sothat

Consider

(4.2,9)

Now from (4.2,4),

and the moment of df about Cis

dG=rXdf

so that from (4.2,7)

dGE = rEdfE = rEvedm

Equation (4.2,9) then becomes

(4.2,10)

ddGE = dt (dbE) - (vE - ~E)vEdm

Since v X v = 0, (4.2,11) becomes

(4.2,11)

(4.2,12)

Page 106: Stability Control EtkinReid

96 Chapter 4. General Equations of Unsteady Motion

Equation (4.2,12) is now integrated as was (4.2,7), and becomes

f dGE = ~ f dbE + "9"E f vEdm

By an argument similar to that for fdf, fdG is shown to be the resultant externalmoment about C, denoted G. f db is called the moment of momentum, or angular mo-mentum of the airplane and is denoted h. Formulas for h are derived in Sec. 4.3. Us-ing (4.2,6) and noting that V X V = 0, (4.2,13) reduces to

(4.2,13)

(4.2,14)

where

The reader should note that, in (4.2,14), both G and h are referred to a movingpoint, the mass center. For a moving reference point other than the mass center, theequation does not in general apply. The reader should also note that (4.2,8) and(4.2,14) are both valid when there is relative motion of parts of the airplane.

The two vector equations of motion of the airplane, equivalent to six scalar equa-tions, are (4.2,8) and (4.2,14)

r, = mVE

GE = hEWhen the wind vector W is not zero, the velocity VEin (4.2,5) is that of the CG rela-tive to FE' The angular momentum h is the same whether W is zero or not (see Exer-cise 4.1). Hence in the general case when wind is present the equations of motionare:

'EfE = mVE

GE = hE"

(4.2,15)

4.3 Evaluation of the Angular Momentum h

We shall want the angular momentum components in FB• Now

h=fdb=frxVdm

So in FB we have

(4.3,1)

where-zox

..

Page 107: Stability Control EtkinReid

4.3 Evaluation of the Angular Momentum h 97

Let the angular velocity of the airplane relative to inertial space' be

WB = [p q rVwhere p, q, r are the rates of roll, pitch and yaw respectively (see Fig. 1.6).

Now the velocity of a point in a rigid rotating body is given by2

(a)where (4.3,2)

-rop

(b)

so that

= (J i'B dm) VB + J i'BiiJBrB dm

The first integral in (4.3,3) vanishes since the origin of r is the eG. When the triplematrix product of the second integral is expanded (see Exercise 4.2) we get the resultfor hg:

(4.3,3)

(4.3,4)

where

(4.3,5)

and

i, = J (y2 + r)dm; t,= J (~ + z2)dm; t,= J (~ + y2)dm

(4.3,6)

t; = Iyx = J xydm; Ixz = t.; = J xz dm; Iyz = I zy = J yz dm

IB is the inertia matrix, its elements being the moments and products of inertia of theairplane. When the xz plane is a plane of symmetry, which is the usual assumption,then

Ixy = Iyz = 0

and the only off-diagonal term remaining is Izr- If the direction of the x axis is so cho-sen that this product of inertia also vanishes, which is always possible in principle,then the axes are principal axes.

1Since there is no need to use the angular velocity relative to any other frame of reference the distin-guishing superscript E is not needed on w.

2See Appendix A.6.

Page 108: Stability Control EtkinReid

98 Chapter 4. General Equations of Unsteady Motion

4.4 Orientation and Position of the AirplaneThe position and orientation of the airplane are given relative to the Earth-fixedframe FE' The CG has position vector r, (see Fig. 4.1), with coordinates (xE, YE' ZE)'

The orientation of the airplane is given by a series of three consecutive rotations,the Euler angles, whose order is important. The airplane is imagined first to be ori-ented so that its axes are parallel to those of FE, (see Fig. 4.2). It is then in the posi-tion CXtYtZt. The following rotations are then applied.

••

1. A rotation '\II' about OZt, carrying the axes to CX:2.Y2Z2 (bringing Cx to its finalazimuth).

2. A rotation 0 about OY2' carrying the axes to CX3Y3Z3 (bringing Cx to its finalelevation).

3. A rotation cI» about OX3' carrying the axes to their final position Cxyz (givingthe final angle of bank to the wings).

In order to avoid ambiguities which can otherwise result in the set of angles (.p,0, c/» the ranges are limited to

~., ..x,

Flight path

XE

YE

o

Earth·fixed axes. FE

Vertical

Figure 4.2 Airplane orientation.

Page 109: Stability Control EtkinReid

4.4 Orientation and Position of the Airplane 99

or7T 7T

--:s():s-2 2-7T:S c/J< 7T or O:s c/J:s 27T

The Euler angles are then unique for most orientations of the vehicle. It should benoted that in a continuous steady rotation, such as rolling, the time variation of c/Jforexample is a discontinuous sawtooth function, and that another exception occurs in avertical climb or dive, when () = ±7T/2. For then (1/1, (), c/J) = (1/1 + a, ±7T/2, -a)gives the same final orientation regardless of the value of a. The above difficultiescan be avoided by using direction cosines" or quaternions" to define the orientation ofthe airplane instead of Euler angles. We use the Euler angles because they give amore physical picture of the airplane attitude than the other alternatives. For a morecomplete discussion of methods of describing vehicle orientation the reader is re-ferred to Hughes (1986).

THE FLIGHT PATH

To track the flight path relative to FE' we need the velocity components in the direc-tions of the axes of FE' These we get by transforming the velocity vector V~ into V~as shown in Appendix A.4.5

(4.4,1)

Here LEB is the matrix of direction cosines that corresponds to the reverse of the se-quence of rotations given above, which are for a transformation from FE to F B' Thus

(4.4,2)

where L; Ly, L, are respectively Lt, ~, L3 of Appendix A.4. Using the rotation matricesgiven in Appendix A.4, and carrying out the multiplication, we get the final result (4.4,3).

[

COS () cos 1/1LEB = cos ()sin 1/1

-sin ()

sin c/Jsin ()cos 1/1 - cos c/Jsin 1/1sin c/Jsin ()sin 1/1 + cos c/Jcos 1/1

sin c/Jcos ()

cos c/Jsin ()cos 1/1 + sin c/Jsin 1/1]cos c/Jsin ()sin 1/1 - sin c/Jcos 1/1

cos c/Jcos ()

(4.4,3)

The differential equations for the coordinates of the flight path are then

[~;] ~ LEBV~

The position of the vehicle CG is obtained by integrating the preceding equation.

3Sec. 5.2 ofEtkin (1972)."Appendix E of ANSIIAIAA (1992).sRecall that the superscript signifies velocity relative to FE' and that the subscript identifies the ref-

erence frame in which the components are given.

Page 110: Stability Control EtkinReid

100 Chapter 4. General Equations of Unsteady Motion

ORIENTATION OF THE AIRPLANE

We now need a set of differential equations from which the Euler angles can be cal-culated. These are obtained as follows: Let (i, j, k) be unit vectors, with subscripts 1,2,3 denoting directions (Xl' YI, z.), and so on of Fig. 4.2.

Let the airplane experience, in time ~t, an infinitesimal rotation from the positiondefined by '1',0, ct>to that corresponding to ('I' + ~'I'), (0 + ~0), (ct> + ~ct». Thevector representing this rotation is approximately

~n ~ i3 ~ct> + j2 ~0 + k, ~'I'and the angular velocity is exactly

(4.4,4)

By using (A 1,4), (A4,3a) and (A4,1O) the unit vectors are found to be in frame FB

[1] [0] [-sin () ]i3B = 0 j2B = co~ i/J klB = cos ()sin i/Jo - sm i/J cos ()cos i/J

(4.4,5)

Recalling that lJJB = [p q r]T, we get

m~R[:] (a)

where (4.4,6)

R~G 0 -sillO ]cos i/J sin i/J cos () (b)-sin c/> cos i/J cos () ,

Inverting (4.4,6) we get the Euler angle rates as

[:] ~T[;] (a)

where (4.4,7)

T~[~ sin i/J tan () CQS 4> tan 0]cos i/J -sin i/J (b)

sin i/J sec () cos i/J sec ()

4.5 Euler's Equations of MotionWe come now to the reason why body axes are so important. Equation (4.2,15) showsthat we need the time derivative of the angular momentum, h, which contains the mo-ments and products of inertia with respect to whatever axes are chosen. If those axesare fixed relative to inertial space, then the inertias will be variables in the equations.This is most undesirable and can be avoided by writing the equations in the frame FB'

in which all the inertias are constant.

Page 111: Stability Control EtkinReid

4.5 Euler's Equations of Motion 101

We begin with the force equation of (4.2,15):'EfE = mVE

Both vectors in (4.5,1) are now expressed in FB components; thus:

(4.5,1)

_ d E _ . E 'ELEJB - m dt (LEBV B) - m(LEBV B + LEBV B)

The derivative of the transformation matrix is obtained from (A.4) as

LEB= LEBtdB

(4.5,2)

(4.5,3)

With (4.5,3), (4.5,2) becomes- E ·ELEBfB = m(LEBCdBVB + LEBVB)

Now premultiply by LBEto get• E - EfB = m(V B + CdBV B)

A similar procedure applied to the moment equation of (4.2,15) leads to

GB = JiB + tdBhB

(4.5,4)

(4.5,5)

The force vector f is the sum of the aerodynamic force A and the gravitational forcemg, that is,

f=mg+A (4.5,6)

where

AB = [X Y Z]T

and

(4.5,7)

We denote V~ = ruE o" wEf and use (4.3,2b) for iiJB. Equations (4.5,4 and 5) arethen expanded using (4.3,4) to yield the desired equations. In doing so we note thatGB = [L M NY and that, because the airplane is assumed to be rigid, t, = O.

X - mg sin (J = m(itE + qwE - ruE) (a)

y + mg cos (J sin 4> = m(iJE + ruE - pwE) (b) (4.5,8)Z + mg cos (J cos 4> = m(w + po" - qw) (c)

L = IxP - Iyiq2 - ,-2) - Ivc(t + pq) - Ixy(q - rp) - (Iy - Iz)qr (a)M = Iyq - Ivc(r - p2) - Ixy(jJ + qr) - Iyif - pq) - (/z - Ix)rp (b) (4.5,9)N = I/- Ixy(p2 - q2) - Iyiq + rp) - Ivc(jJ - qr) - (Ix - Iy)pq (c)

CHOICE OF BODY AXES

The equations derived in the preceding sections are valid for any orthogonal axesfixed in the airplane, with origin at the eG, and known as body axes. Since most air-craft are very nearly symmetrical, it is usual to assume exact symmetry, and to letCxz be the plane of symmetry. Then Cx points "forward," Cz "downward," and Cy to

Page 112: Stability Control EtkinReid

102 Chapter 4. General Equations of Unsteady Motion

ZpPrincipal axes Stability &Xes

E>OFigure 4.3 Illustrating two choices of body axes.

the right. In this case, the two products of inertia, Ixy and IyZ' are zero, and (4.5,9) areconsequently simplified.

The directions of Cx and Cz in the plane of symmetry are conventionally fixed inone of three ways (see Fig. 4.3).

..

Principal Axes

These are chosen to coincide with the principal axes of the vehicle, so that the re-maining product of inertia Izx vanishes; (4.3,4 and 5) then yield

b, = IxP

h, = LIlh, = Izr

(4.5,10)

Stability AxesThese are chosen so that Cx is aligned with V in a reference condition of steady

symmetric flight. In this case, the reference values of v and w are zero, and the axesare termed stability axes. These axes are commonly used, owing to the simplifica-tions that result in the equations of motion, and in the expressions for the aerody-namic forces.

With this choice, it should be noted that for different initial flight conditions theaxes are differently oriented in the airplane, and hence the values of Ix, Iz' and Izx willvary from problem to problem. The "stability axes," just as the principal axes, arebody axes that remain fixed to the airplane during the motion considered in anyoneproblem.

The following formulas are convenient for computing Ix, Iz' Izx when the valuesIx and I, are known for principal axes (see Exercise 4.4).

" "Ix = Ixp cos" € + Izp sirr' €

I,= Ixp sirr' € + Izp cos" € (4.5,11)1

t.; = 2" (/zp - Ixp) sin 2€

where e = angle between xp (principal axis) and Xs (stability axis), positive as shown(see Fig. 4.3).

•••

Page 113: Stability Control EtkinReid

4.7 The Equations Collected 103

Body AxesWhen the axes are neither principal axes nor stability axes, they are usually

called simply body axes. In this case the x axis is usually fixed to a longitudinal refer-ence line in the airplane. These axes may be the most convenient ones to use if theaerodynamic data have been measured by a wind-tunnel balance that resolves theforces and moments into body-fixed axes rather than tunnel-fixed axes.

4.6 Effect of Spinning Rotors on the Euler EquationsIn evaluating the angular momentum h (see Sec. 4.3) it was tacitly assumed that theairplane is a single rigid body. This is implied in the equation for the velocity of anelement (4.3,2). Let us now imagine that some portions of the airplane mass are spin-ning relative to the body axes; for example, rotors of jet engines, or propellers. Eachsuch rotor has an angular momentum relative to the body axes. This can be computedfrom (4.3,4) by interpreting the moments and products of inertia therein as those ofthe rotor with respect to axes parallel to Cxyz and origin at the rotor mass center. Theangular velocities in (4.3,4) are interpreted as those of the rotor relative to the air-plane body axes. Let the resultant relative angular momentum of all rotors be h', withcomponents (h~,h;, h~)in FB' which are assumed to be constant. It can be shown thatthe total angular momentum of an airplane with spinning rotors is obtained simply byadding h' to the h previously obtained in Sec. 4.3 (see Exercise 4.5). The equationthat corresponds to (4.3,4) is then?

(4.6,1)

Because of the additional terms in the angular momentum, certain extra terms appearin the right-hand side of the moment equations, (4.5,9). Those additional terms,known as the gyroscopic couples, are

In the L equation:In the M equation:In the N equation:

qh~- rh';

rh~ - ph~ph; - qh~

(4.6,2)

As an example, suppose the rotor axes are parallel to Cx, with angular momentumh' = un. Then the gyroscopic terms in the three equations are, respectively, 0, Inr,and -Inq.

4.7 The Equations CollectedThe kinematical and dynamical equations derived in the foregoing are collected be-low for convenience. The assumption that Cxz is a plane of symmetry is used, so thatIxy = Iyz = 0, and (4.6,2) are added to (4.5,9) to give (4.7,2).

6Note that the inertias of the rotors are also included in lB'

Page 114: Stability Control EtkinReid

104 Chapter 4. General Equations of Unsteady Motion

X - mg sin 0 = m(uE + qW' - ruE) (a)Y + mg cos 8 sin c/>= m( iJE + ruE - pwE) (b) (4.7,1) •Z + mg cos 0 cos c/>= m(wE + pvE - quE) (c)

L = I,.jJ - Izxt + qnl, - Iy) - Izxpq + qh~ - rh~ (a)M = IiJ + rptl; - Iz) + Izx(p2 - r2) + rh~ - ph~. (b) (4.7,2)N = Izt - Izxp + pq(Iy - Ix) + I~r + ph'; - qh~ (c)

p=<b-,psinO (a)q = (j cos c/>+ ,p cos 0 sin c/> (b)r = ,p cos 0 cos c/>- (j sin c/> (c)

<b= p + (q sin c/>+ rcos c/» tan 8 (d) (4.7,3)(j = q cos c/>- r sin c/> (e),p = (q sin c/>+ r cos c/» sec 0 (f)

XE = uE cos 8 cos l/J + vE(sin c/>sin 0 cos l/J - cos c/>sin l/J)+ W'(cos c/>sin 0 cos l/J + sin c/>sin l/J) (a)

YE = uE cos 0 sin l/J + vE(sin c/>sin 0 sin l/J + cos c/>cos l/J) (4.7,4)

+ W'(cos c/>sin 0 sin l/J - sin c/>cos l/J) (b)ZE = -uE sin 8 + vE sin c/>cos 0 + W' cos c/>cos 0 (c)

uE=u+Wx (a)vE = v + W (b) (4.7,5)yW'=w+Wz (c)

4.8 Discussion of the EquationsThe above equations are quite general and contain few assumptions. These are:

1. The airplane is a rigid body, which may have attached to it any number ofrigid spinning rotors.

2. Cxz is a plane of mirror symmetry.3. The axes of any spinning rotors are fixed in direction relative to the body

axes, and the rotors have constant angular speed relative to the body axes.

The equations of Sec. 4.7 are many and complex. They consist of 15 couplednonlinear ordinary differential equations in the independent variable t, and three alge-braic equations. Before we can identify the true dependent variables, however, wemust first consider the aerodynamic forces (X,Y,Z) and moments (L,M,N). It is clearthat these must depend in some manner on the relative motion of the airplane with re-spect to the air, given by the linear and angular velocities V and iAJ, on the controlvariables that fix the angles of any moveable surfaces and on the settings of anypropulsion controls that determine the thrust vector. Thus it is universally assumed inflight dynamics that the six forces and moments are functions of the six linear andangular velocities (u,v,w,p,q,r) and of a control vector. The latter clearly depends to

Page 115: Stability Control EtkinReid

4.8 Discussion of the Equations 105

some extent on the particular airplane, but we can, with adequate generality, write itas

c = [Sa Se Sr sptof which the first three are the familiar aileron, elevator, and rudder angles, and thelast is the throttle control. Other components can be added to the control vector asneeded to meet special requirements-for example, direct lift control. The controlvariables, from the standpoint of the mathematical system, are arbitrary functions oftime. How they are determined is the subject of later sections. The wind vector whosecomponents appear in (4.7,5) would ordinarily be a known function of position r,with its components given in frame FE' Its components in F B are WB = LBEWE'

The true implicit dependent variables of the system are thus seen to be 12 innumber:

CO position: XE, YE' ZE

Attitude: rfJ, 0, 4JVelocity: uE, VE, wE

Angular velocity: p, q, r

Of the 15 differential equations we note that 3 of (4.7,3) are not independent, sothe number of independent equations is actually 12, the same as the number of de-pendent variables. The mathematical system is therefore complete.

A useful view of the equations is given in the block diagram of Fig. 4.4, which isspecialized to the case of zero wind. Each block represents one set of equations, withinputs and outputs (the dependent variables). All the inputs needed for the left-handside are generated as outputs on the right, except for the control inputs, which remainto be specified. The nature of the mathematical problem that ensues is very much

B,q, u --[>---uu,v,w iJ --[>---vp,q,r

Control forces W --[>---w

u,v,w p --[>---pp,q,r q --[>---q

Control moments r --[>---r

J,B,q, --[>---q,; ---D>--B

s.s.r ~ --IT>--l/I

u, v, W xE ---D>-- ~E

YB --[>--- Ys

q" B, l/I zB ---D>-- zB

Figure 4.4 Block diagram of equations for vehicle with plane of symmetry. Body axes. Flat-Earthapproximation. No wind.

Page 116: Stability Control EtkinReid

106 Chapter 4. General Equations of Unsteady Motion

governed by the specifics of the control inputs. In the following paragraphs, we dis-cuss the various cases that commonly occur in engineering practise and in research.

STABILITY PROBLEMS-CONTROLS FIXED

In these problems the airplane is considered to be disturbed from an initially steadyflight condition, with the controls locked in position. Thus, c is zero or a known con-stant. The equations are nonlinear and, consequently, extremely difficult to deal withanalytically. In many problems of practical importance, it is satisfactory to linearizethe equations by dealing only with small perturbations from the reference condition.In that case we obtain a set of homogeneous linear differential equations with con-stant coefficients, a type that is readily solved. Problems of this class are treated inChap. 6.

STABILITY PROBLEMS-CONTROLS FREE

The free-control case is of interest primarily only for manually controlled airplanes.In that case, one or more of the primary control systems is presumed to be freed as in"hands off" by the pilot. The variation of the control angles with time, which is ofcourse needed for the aerodynamic force and moment inputs, is then the result of aninteraction between the dynamics and aerodynamics of the vehicle and those of thecontrol system itself, which is usually simplified as a system with one degree of free-dom relative to FB• For a derivation of these equations see Etkin (1972, Sec. 11.3).Each such free control adds one dependent variable and one equation to the mathe-matical system.

STABILITY PROBLEMS-AUTOMATIC CONTROLS

When the airplane is under the control of an APeS the controls are neither fixed norfree. The control vector c, which fixes the control inputs of Fig. 4.4, are then deter-mined by the feedback loop that activates the control systems in response to the val-ues of the 12 dependent variables and inputs from other sources such as navigation,guidance, or fire control systems, and so on. Problems of this class are studied inChap. 8.

RESPONSE TO CONTROLS

The effectiveness of an airplane's controls is conventionally studied by specifying thevariation of (Se' Sr' Sa' Sp) with time arbitrarily, e.g., a step-function input of aileronangle. The airplane equations of motion then become inhomogeneous equations for(u, D, w), (p, q, r), (0, cP, r/J) with the control angles as forcing functions. Problems ofthis type are treated in Chap. 7. '.RESPONSE TO ATMOSPHERIC TURBULENCE

The motion of an airplane, and the forces that act on it, as a consequence of the tur-bulent motion of the atmosphere, are very important for both design and operation.The associated mathematical problems are treated with the same general equations as

.•.

Page 117: Stability Control EtkinReid

4.9 The SmaU-Disturbance Theory 107

given above. (uE, VE, wE) then have to be sums of (u, v, w) and the velocity of the at-mosphere at the CG, and additional complications arise from the fact that the relativewind varies, in general, from point to point on the airplane. This case is treated indepth in Chap. 13 of Etkin (1972) and in Etkin (1981).

INVERSE PROBLEMS

A class of problems that has not received much attention in the past, but that is never-theless both interesting and useful, is that in which some of the 12 variables usuallyregarded as dependent are prescribed in advance as functions of time. An equal num-ber of equations must then be dropped in order to maintain a complete system. Thisis the kind of problem that occurs when we ask questions of the type "Given the air-plane motion, what pilot action is required to produce it?" Such questions may be rel-evant to problems of control design and maneuvering loads.

The mathematical problem that results is generally simpler than those of stabilityand control. The equations to be solved are sometimes algebraic, sometimes differen-tial. A decided advantage is the ability of this approach to cope with the nonlinearequations of large disturbances.

Another category is the mathematical problem that arises in flight testing whentime records are available of some control variables and some of the 12 dependentvariables. The question then is "What must the airplane parameters be to produce themeasured response from the measured input?" (See Etkin, 1959, Chap. 11; AGARD1991; Maine and Iliff, 1986.) This is an example of the important "plant identifica-tion" problem of system theory.

4.9 The Small-Disturbance TheoryAs remarked in Sec. 4.8, the equations of motion are frequently linearized for use instability and control analysis. It is assumed that the motion of the airplane consists ofsmall deviations from a reference condition of steady flight. The use of the small-dis-turbance theory has been found in practice to give good results. It predicts with satis-factory precision the stability of unaccelerated flight, and it can be used, with suffi-cient accuracy for engineering purposes, for response calculations where thedisturbances are not infinitesimal. There are, of course, limitations to the theory. It isnot suitable for solutions of problems in which large disturbance angles occur, for ex-ample <I>~ 1T12.

The reasons for the success of the method are twofold: (1) In many cases, themajor aerodynamic effects are nearly linear functions of the disturbances, and (2) dis-turbed flight of considerable violence can occur with quite small values of the linear-and angular-velocity disturbances.

NOTATION FOR SMALL DISTURBANCES

The reference values of all the variables are denoted by a subscript zero, and thesmall perturbations by prefix a.When the reference value is zero, the a may be omit-ted. All the disturbance quantities and their derivatives are assumed to be small, sothat their squares and products are negligible compared to first-order quantities.

Page 118: Stability Control EtkinReid

108 Chapter 4. General Equations of Unsteady Motion

The reference flight condition is assumed for convenience to be symmetric andwith no angular velocity. Thus Vo = Po = qo = '0 = 4>0= rfio= O.Furthermore, sta-bility axes are selected as standard in this book, and thus Wo = 0 for all the problemsconsidered. Uo is then equal to the reference flight speed, and ()o to the reference an-gle of climb (not assumed to be small). In dealing with the trigonometric functions inthe equations the following relations are used:

sin (00 + !i.0) = sin 00 cos !i.0 + cos ()o sin !i.0

* sin 00 + !i.0 cos 00

cos (00 + !i.0) = cos 00 cos !i.0 - sin 00 sin !i.0* cos ()o - !i.0 sin 00

FURTHER ASSUMPTIONS

(4.9,1)

The small-disturbance equations will be slightly restricted by the adoption of twomore assumptions, which correspond to current practice. These are

1. The effects of spinning rotors are negligible. This is the case when the air-plane is in gliding flight with power off, when the symmetrical engines haveopposite rotation, or when the rotor angular momentum is small.

2. The wind velocity is zero, so that VE = V

LINEARIZATION

When the small-disturbance notation is introduced into the equations of Sec. 4.7, theadditional assumptions noted above are incorporated, and only the first-order terms indisturbance quantities are kept, then the following linear equations are obtained.

Xo + ax - mg(sin 00 + !i.0 cos 00) = m!i.uYo + !i.Y + mg4>cos 00 = m(iJ + uo')

20 + !i.Z + mg(cos 00 - !i.0 sin 00) = mew - u~)

La + tiL = IxP - IzxtMo + tiM = 1/1.

No + !iN = -Izxp + Ii!i.e = q

ei> = P + r tan 00, P = ei> - rfr sin ()o

rfr = r sec 00

xE = (uo + !i.u) cos 00 - uo!i.Osin 00 + w sin 00

YE = uorficos 00 + V

i.E = -(uo + !i.u) sin 00 - uo!i.Ocos 00 + w cos 00

(a)

(b) (4.9,2)

(c)

(a)

(b) (4.9,3)

(c)

(a)(b) (4.9,4)

(c)

(a)(b) (4.9,5)

(c)

REFERENCE STEADY STATEIf all the disturbance quantities are set equal to zero in the foregoing equations, thenthey apply to the reference flight condition. When this is done, we get the following

Page 119: Stability Control EtkinReid

4.9 The Small-Disturbance Theory 109

relations, which may be used to eliminate all the reference forces and moments fromthe equations:

Xo - mg sin 00 = 0

Yo = 0Zo + mg cos 00 = 0La =Mo =No = 0

YEo = 0,

(4.9,6)

We further postulate that in the reference steady state the aileron and rudder anglesare zero. When (4.9,6) are substituted into (4.9,2 to 4.9,5), (4.9,3) are solved for pand i, and the equations are rearranged, the result is (4.9,7 to 4.9,10).

sxAu = - - gAO cos 00m

AYiJ = - + gcf> cos 00 - uor

m

AZW = - - gAO sin 00 + UofJ.

m

p = (Ixlz - I~)-l (/ilL + IxzAN>

AMq=-t,

i = (Ixlz - t;)-l(/fl + IxAN>AO = q

4> = p + r tan °0; p = 4> - ~ sin 00

~ = rsec ()o

AXE = Au cos 00 - uoAO sin 00 + w sin 00

AYE = uor/J cos 00 + V

AZE = - Au sin 00 - uoA °cos 00 + w cos 00

THE LINEAR AIR REACTIONS

(a)

(b) (4.9,7)

(c)

(a)

(b) (4.9,8)

(c)

(a)(b) (4.9,9)

(c)(a)

(b) (4.9,10)

(c)

At the heart of the subject of atmospheric flight mechanics lies the problem of deter-mining and describing the aerodynamic forces and moments that act on a given bodyin arbitrary motion. It is primarily this aerodynamic ingredient that distinguishes itfrom other branches of mechanics. Aerodynamic forces and moments are strictlyspeaking functionals of the state variables. Consider for example the time-dependentlift L(t) on a wing with variable angle of attack a(t). Because the wing leaves behindit a vortex wake that in general generates an induced velocity field at the wing, andbecause hysteresis is present in flow separation processes, the aerodynamic field thatfixes the lift at any given moment is actually dependent not only on the instantaneousvalue of a but strictly speaking on its entire past history. This functional relation isexpressed by

c

Page 120: Stability Control EtkinReid

110 Chapter 4. General Equations of Unsteady Motion

L(t) = L[a( T)] -00 S TS t (4.9,11)

When a( T) can be expressed as a convergent Taylor series around t, i.e.

a(T) = a(t) + (T - t)a(t) + i(T - t)2ii(t) + '" (4.9,12)

then the infinite series a(t), a(t), ii(t) ... can replace a(T) in (4.9,11). For time t = tothis becomes

L(to) = Lia, a, ii ... ) (4.9,13)

where a, a ... are values at time to. Thus the lift at time to is in this case determinedby a and all its derivatives at time to' A further series expansion of the right-hand sideof (4.9,13) around a(to), a(to), etc. yields

tlL(to) = LcAa + iLaa(~af + ... + LOt~a + ll-OtOt(~af +... (4.9,14)

in which all the products and powers of ~a, Aa .. , appear, and where

La = ~L I etc. (4.9,15)ua "'oCto)

The classical assumption of linear aerodynamic theory, due to Bryan (1911) is to ac-cept the linear reduction of (4.9,14) as a representation of the aerodynamic force,even when Aa(t) is not an analytic function as implied by (4.9,12), i.e.

AL(to) = LcAa + LaAa + LaAii + ... (4.9,16)

Derivatives such as La in (4.9,16) are known as the stability derivatives, or more gen-erally as aerodynamic derivatives. For most forces and state variables, only the firstterm of (4.9,16) is kept, but in some cases, terms up to the second derivative must beretained for sufficient accuracy. This assumption has been found to work extremelywell over a wide range of practical applications. Occasionally the addition of nonlin-ear terms such as iLaa(Aa)2 == L.",(Aaf can extend the. useful range considerably.Another way of including nonlinear effects is to treat the derivatives as functions ofthe variables, for example, La = La(a).

A major fraction of the total effort in aerodynamic research in the past has beendevoted to the determination, by theoretical and experimental means, of the aerody-namic derivatives needed for application to flight mechanics. A great mass of infor-mation about these parameters has now been accumulated and Chap. 5 is devoted tothis topic.

For a truly symmetric configuration, it is evident that the side force Y, the rollingmoment L, and the yawing moment N will all be exactly zero in any condition ofsymmetric flight, that is, when the plane of symmetry remains in a fixed verticalplane. In that case, {3,p, r, 4>, tfJ, and YE are all identically zero. Thus the derivativesof the asymmetric or lateral forces and moments, Y,L, N with respect to the symmet-ric or longitudinal motion variables u, w, q are zero. In writing out the complete lin-ear expression for the aerodynamic forces and moments, we use this fact, and in ad-dition make the further approximations:

1. We may neglect as well all the derivatives of the symmetric forces and mo-ments with respect to the asymmetric motion variables.

2. We may neglect all derivatives with respect to rates of change of motion vari-ables except for Zw and Mw.

Page 121: Stability Control EtkinReid

4.9 The Small-Disturbance Theory 111

3. The derivative Xq is also negligibly small.4. The density of the atmosphere is assumed not to vary with altitude (see Sec.

6.5).

It should be emphasized that none of these assumptions is basically necessary forthe solution of airplane dynamics problems. They are made as a matter of experienceand convenience. When it appears necessary to do so, any of the terms dropped canbe restored into the equations. With these assumptions, however, the linear forces andmoments are:

AX = Xudu + Xww + AXedY = Yvv + Ypp + Y,.r+ dYedZ = Zudu + Zww + Zww + Zqq + dZeM = Lvv + Lpp + L,.r + MeliM = Mudu + Mww + Mww + Mqq + liMeM=Nvv +Npp + N,.r + Me

(a)(b)

(c)

(d)(e)

(f)

(4.9,17)

In the above equations, the terms on the right with subscript c are control forcesand moments that result from the control vector c. Explicit forms for the controls willbe introduced as they are needed in the following.

Aerodynamic Transfer FunctionsThe preceding equations are subject to the theoretical objection (not of great

practical importance) that the Bryan formulation for the aerodynamics is not quitesound even within the restriction of linearity. This is readily illustrated by consideringthe lift on a wing following a step change in angle of attack. Let da be given byda(t) = aol(t) where ao is a constant. For t > 0, the Bryan formula (4.9,16) gives

M(t) = L",ao = const

whereas in fact the lift undergoes a transient approach to the asymptote L",ao, the de-tails of which depend on the wing shape and the Mach number. Equation (4.9,16)fails in this case because da is not an analytic function, having a discontinuity att = O.Now the transient process is often well approximated as a linear one, and assuch is subject to exact representation by linear mathematics in the form of an indi-cial function (Tobak, 1954), or an aerodynamic transfer function (Etkin, 1956). Theimplementation of these alternative representations of aerodynamic force is describedin Etkin (1972, Sec. 5.11).

THE LINEAR EQUATIONS OF MOTION

When (4.9,17) are substituted into (4.9,7 to 4.9,10) two of the equations [(4.9,7c) and(4.9,8b)], contain w terms on the right-hand side. In order to retain the desired formwith first derivatives of the dependent variables on the left, we have to solve thesetwo equations simultaneously for w and if. The result is presented in matrix form in(4.9,18 and 4.9,19). Here the equations are divided into two groups, termed longitu-dinal and lateral, for reasons that are explained as follows.

Page 122: Stability Control EtkinReid

Longitudinal Equations, Eq. (4.9,18):

Auw

=qAiJ

o

IIII

m ,m: I--------------------~-------------------~------------------------~----------------z; : z, : z, + mu; : -mg sin e,1 I I

m-~ : m-~ : m-~ : m-~____________________ L L L _

1 [ MwZu ] i 1 [ u.z; ]: 1 [ Mw(Zq + muo) ] i u; mg sin «- Mu + I - Mw + I - M + I - ------t, (m -z,» : t, (m -Zw) : t, q (m -Zw) : Iy(m -Zw)--------------------r-------------------r------------------------r----------------1 I 1

: 0 : 1 :I 1 I

o -g cos (Jo

oAXE = Au cos (Jo + w sin (Jo - UO A(J sin (Jo

AtE = -Au sin (Jo + w cos (Jo - Uo A(Jcos (Jo

AuwqA(J

AXem

AZe

m-~+ -----------------------We u; AZe--+----t, t,

o

Page 123: Stability Control EtkinReid

Lateral Equations, Eq. (4.9,19):

p

, ,Yv Yp

: (Yr_ u) ! g cos 00m m: m 0 ,--------------~-------------~------------i---------

(Lv + I'zxN,) : (Lp + I'zxN.) : (Lr + I'dv.) i 0I~ v:I~ P:I~ 'l :

= --------------~-------------~------------~---------

( N '( N)'( N)'1'1 +_v): 1'1 +--...£.. : 1'1+_r : 0zxL'v I' I zxL'p l' ' zxL'r I' ,

Z I Z I Z I--------------~-------------~------------1---------, I ,

o : 1 : tan 00 : 0, , ,

t

r;,= r sec 00

J1.YE = uol/J cos 00 + V

I~ = (IJz - Iz})/lzI~= (IJz - Iz})/lxI~ = Iz/(lJz - Ix/)

V

p

r

m

M__ C + I'dv.I~ C

+ ---------------MI~ + __c

C I~

o

Page 124: Stability Control EtkinReid

114 Chapter 4. General Equations of Unsteady Motion

As a consequence of the simplifying assumptions made in their derivation, the pre-ceding equations have the exceedingly useful property of splitting into two indepen-dent groups. Suppose that <P, v, p, r, aYe, aLe and !:1Neare identically zero. Then(4.9,19) are all identically satisfied. The remaining equations (4.9,18) form a com-plete set for the six homogeneous variables au, w, q, a(), Ii.xE, aZE' Thus we mayconclude that modes of motion are possible in which only these variables differ fromzero. Such motions are called longitudinal or symmetric, and the corresponding equa-tions and variables are likewise named. Conversely, if the longitudinal variables areset equal to zero, the remaining six equations (4.9,19) form a complete set for the de-termination of the variables <p, 1/1, v, p, r, YEo These are known as the lateral variables,the corresponding equations and motions being likewise named.

It is worthwhile recording here the specific assumptions upon which this separa-tion depends. A study of the various steps that have led to the final equations revealsthese facts-the existence of the pure longitudinal motions depends on only two as-sumptions:

1. The existence of a plane of symmetry.2. The absence of rotor gyroscopic effects.

The existence of the pure lateral motions, however, depends on more restrictive ap-proximations; namely

1. The linearization of the equations.2. The absence of rotor gyroscopic effects.3. The neglect of all aerodynamic cross-coupling (approximation 1 p. 110).

If the equations were not linearized, then there would be inertial cross-couplingbetween the longitudinal and lateral modes, as evidenced by terms such as mpv" in(4.7,lc) and rp(IIC - /z) in (4.7,2b). That is, motion in the lateral modes would inducelongitudinal motion but not the reverse.

Equations (4.9,18 and 4.9,19) are both in the desired first-order form, commonlyreferred to as state vector form, conventionally written in vector/matrix notation as

i=Ax+Bc (4.9,20)

Here x is the state vector, c is the control vector, and A and B are system matrices.The state vectors for the longitudinal and lateral systems are, respectively:

x = [au w q a()t

x = [v p r <Pt (4.9,21)

and the matrices A for the two cases can be inferred from the full equations. The de-pendent variables XE, YE' ZE and 1/1 are not included in the state vectors because theydo not appear on the right-hand side of the equations. The matrix B will be discussedlater when we come to the analysis of controlled motions.

Page 125: Stability Control EtkinReid

4.10 Nondimensional System 115

4.10 Nondimensional SystemThe reader will already be familiar with the great advantage of using nondimensionalcoefficients for aerodynamic forces and moments such as lift, drag, and pitching mo-ment. In this way the major effects of speed, size, and air density are automaticallyaccounted for. Similarly we need nondimensional coefficients for the many deriva-tives-Xu and so on-that occur in (4.9,18 and 4.9,19). Unfortunately, there is nouniversally accepted standard for these coefficients, although attempts have beenmade to devise one (e.g., ANSIIAIAA, 1992). The student, and indeed the practisingengineer, should be sure to note carefully the exact notation and definitions employedin any reference material or data sources being used. The notation and definitionsused in this book are essentially the NASA system, which is widely used.

Before presenting this system, we digress briefly to a dimensional analysis of thegeneral flight dynamics problem. This helps to provide insight into what the true un-derlying variables are, and provides a basis for what follows. Imagine a class of geo-metrically similar airplanes of various sizes and masses in steady unaccelerated flightat various heights and speeds. Suppose that one of these airplanes is subjected to adisturbance. After the disturbance, some typical nondimensional variable 71' varieswith time. For example, 71' may be the angle of yaw, the load factor, or the helix anglein roll. Thus, for this one airplane, under one particular set of conditions we shallhave

71' = f(t) (4.10,1)

Let it be assumed that this equation can be generalized to cover the whole class ofairplanes, under all flight conditions. That is, we shall assume that 71' is a function notof t alone, but also of

uo, p, m, /, g, M, RN

where m is the airplane mass and l is a characteristic length. Instead of (4.10,1), then,we write

71' = f(uo, p, m, l, g, M, RN, t) (4.10,2)

Buckingham's 71' theorem (Langhaar, 1951) tells us that, since there are nine quanti-ties in (4.10,2) containing three fundamental dimensions, L, M, and T, then there are9 - 3 = 6 independent dimensionless combinations of the nine quantities. These sixso-called 71' functions are to be regarded as the meaningful physical variables of theequation, instead of the original nine. Two systems of the same class are dynamicallysimilar when all the 71' functions of one are numerically equal to those of the other.By inspection, we can easily form the following six independent nondimensionalcombinations:

m uot u~71', M, RN, pP' / ' t«

Following the 71' theorem, we write as the symbolic solution to our problem

(m uot U~)

71'= f M,RN, pP' -/-, Ii (4.10,3)

Page 126: Stability Control EtkinReid

116 Chapter 4. General Equations of Unsteady Motion

The effects of the six variables m, p, 1, g, uo, and t are thus seen to be compressed intothe three combinations: m/pP, u't/lg, and uotll. We replace 13 by Sl, where S is a char-acteristic area, without changing its dimensions, and denote the resulting nondimen-sional quantity m/ pSI by 11-. The quantity l/uo has the dimensions of time and is de-noted t*. The quantity u~/lg is the Froude number (FN). Equation 4.10,3 thenbecomes

1T' = f(M, RN, FN, 11-, t1t*) (4.10,4)

The significance of (4.10,4) is that it shows rr to be a function of only five vari-ables, instead of the original eight. The result is of sufficient importance that it is cus-tomary to elaborate on it still further. Since II- is the ratio of the airplane mass to themass of a volume Sl of air, it is called the relative mass parameter or relative densityparameter. It is smallest at sea level and increases with altitude.

The main symbols for which nondimensional forms are wanted are listed inTable 4.1. The nondimensional item in column 3 is obtained by dividing the corre-sponding dimensional item of column 1 by the divisor in column 2. In the small-dis-turbance case, () and w are aerodynamic angles, for then

V = [(uo + au)2 + v2 + W2)]l/2

From (1.6,4)

ax = tan-1(w/u) = tan-1[w/(uo + au)]

To first order in v, w, and au these are

V == Uo + au

Wax== - =w

Uo

Table4.lThe Nondimensional System

(1) (2) (3) t

DimensionalQuantity

Divisor-General

Case

Divisor-Small-

DisturbanceCase

NondimensionalQuantity

X,Y,ZmgM

L,N

~py2s!py2sipV2Seipy2Sb

Vzvt:2VlbpSe/2

pS(C/2)3pS(bI2)3

ipv2sipy2Sipy2SeipV2Sb

Uo2uJe2uJbpSe/2

pS(e/2)3pS(bI2)3

t* = e/(2Uo)

c; c; c,CwCmc;c,

it, D, W (1)&,q

ffi,p,fJ.tj

A "Y At; t; Izx1

u,V,W

a,q/3,p, r

mt,

t; t; t.;t

Note: (1) f3 and ax are used interchangeably with D and lV,respectively, in the small disturbance case.

Page 127: Stability Control EtkinReid

4.10 Nondimensional System 117

Table 4.2Longitudinal NondimensionaJ Derivatives

Uaq11

and similarly

v V(3= sin-1 - == - = 1)

V Uo

NONDIMENSIONAL STABILITY DERIVATIVES

The nondimensional stability derivatives are the partial derivatives of the force andmoment coefficients in lines 1, 3, and 4 of Table 4.1 with respect to the nondimen-sional motion variables in lines 5, 6, and 7. The notation for these is displayed in Ta-bles 4.2 and 4.3. Each entry in the tables represents the derivative of the columnheading with respect to the row variable.

Since ax differs from a only by a constant (the angle between the zero-lift lineand the x axis), then Aax = Aa, a/dax == aIda, and no distinction need be made be-tween these two derivatives.

NONDIMENSIONAL EQUATIONS

J

It is possible with the definitions given in Tables 4.1;-4.3 to make the equations ofmotion entirely nondimensional, and such equations have been widely used in thepast, especially for analytical work (see Etkin, 1972 and 1982). The prevailing cur-rent practise in design and research, however, is to use the dimensional equations andprogram them for calculation on a digital computer. We are therefore not includingthe nondimensional equations in this book. There is no real loss in so doing however,since any analytical results that are obtained with the dimensional equations can sub-sequently be expressed, for maximum generality, in nondimensional form. Examplesof this are contained in Chaps. 6 and 7.

i

Table 4.3Lateral Nondimensional Derivatives

Page 128: Stability Control EtkinReid

118 Chapter 4. General Equations of Unsteady Motion

Table 4.4Longitudinal Dimensional Derivatives

x z M

uwqw

4.11 Dimensional Stability DerivativesWe now need expressions for the derivatives that appear in (4.9,18) and (4.9,19) interms of the nondimensional derivatives. A few examples of these are derived as fol-lows to illustrate the procedure, and the whole set needed for (4.9,17) is displayed inTables 4.4 and 4.5. Derivatives with respect to iJ or /3 are usually negligible and arenot included.

THE Z DERIVATIVES

From Table 4.1, Z = CzipV2S where V2 = u2 + v2 + w2 and u = Uo + au. Hence

(az) (av) 2 (acz)Z; = au 0 = czopuoS au 0 + ipuo S au 0

where the subscript zero indicates the reference flight condition. But 2V(aVlau) =2u, and hence (aVlau)o = 1.Also

(acz ) 1 (acz) 1au 0 = Uo au 0 = Uo CZu

From (4.9,6)

Zo = -mg cos 80

hence

Czo = -Cwo cos 80

so that

Table 4.5Lateral Dimensional Derivatives

y L N

vpr

!puobSC/j3

\puob2SC/\pUob2SC;'

!PuobSCn/l\puob2SCn\pUob2SC~

Page 129: Stability Control EtkinReid

4.11 Dimensional Stability Derivatives 119

Also, since (dVldW)o = 0, then

_ ( dZ) _ 1 2 (dCz)Zw - dW 0 - "2 pufyS dW 0

But W = uoax, so that

1 (dCz) 1z = - pu {' - = - pu {'Cw 2 (j>J da 2 o- Za

ex 0

In a similar way,

(dZ) 1 2 (dCz) 1 ( sc, )

Zw = dW 0"2 puoS dW 0 = "2 punS dil!x 0

But

Hence

Also,

But

2uo A

q=~qC

Hence

THE X DERIVATIVES

These are found in a manner similar to the Z derivatives. In this instance from (4.9,6)

CX{) = Cwo sin 00

THE M DERIVATIVES

These are also found in a manner similar to the Z derivatives. In this case we startwith M = CmipV2Sc and note from (4.9,6) that Crne = O.

Page 130: Stability Control EtkinReid

120 Chapter 4. General Equations of Unsteady Motion

THE L DERIVATIVES

bFrom Table 4.1, L = ClPV2S 2' Hence

t; = (aL) = pu~ !!.- ( acl) = puoS !!.- ( aCI )au 0 2 au 0 2 a~ 0

b= puoS 2 Clfj

Also

Similarly

_ ( aL) _ 2 b (acl)L - - -puOS- -r ar 0 2 ar 0

1= 4" puob2SCIr

THE N DERIVATIVES

These are found in a manner similar to the L derivatives.

THE Y DERIVATIVES

These are also found in a manner similar to the L derivatives. In this case we startwith Y = Cylpy2S.

4.12 Elastic Degrees of FreedomIn the preceding sections we have presented the "main" equations of motion, that is,those associated with the six rigid-body degrees of freedom. Now it is known that thestability and control characteristics of flight vehicles may be profoundly influencedby the elastic distortions of the structure under aerodynamic load (AGARD, 1970;Milne, 1964; McLaughlin, 1956; Rodden, 1956). Additionally, there are phenomenanot primarily related to stability and control, but rather to structural integrity, inwhich elastic deformation is a primary element-i.e., structural divergence and flut-ter. In order to understand and analyze all these effects, one needs the equations thatgovern the elastic deformations, and as well the changes that such deformation intro-duces into the six main dynamical equations.

A full treatment of this branch of flight mechanics-aeroelasticity and structuralvibration-is beyond the scope of this text, and the reader is referred to (Bispling-hoff, 1962 and Dowell, 1994) for comprehensive treatises on it. Here we content our-selves with presenting the framework of the analysis, but omit most of the structural

Page 131: Stability Control EtkinReid

4.12 Elastic Degrees of Freedom 121

and aerodynamic details. Enough material is given, however, to show how the staticand dynamic deformations are integrated into the preceding mathematical model ofthe "gross" vehicle motion.

The deformation analysis is almost invariably treated by a linear theory, evenwhen the rigid-body motion is not. We shall therefore assume at the outset that thedistortional motions are "small" and that all the associated aerodynamic forces arelinear functions.

THE METHOD OF QUASISTATIC DEFLECTIONS

Many of the important effects of distortion can be accounted for simply by alteringthe aerodynamic derivatives. The assumption is made that the changes in aerody-namic loading take place so slowly that the structure is at all times in static equilib-rium. (This is equivalent to assuming that the natural frequencies of vibration of thestructure are much higher than the frequencies of the rigid-body motions.) Thus achange in load produces a proportional change in the shape of the vehicle, which inturn influences the load. Examples of this kind of analysis are given in Sec. 3.5 (ef-fect of fuselage bending on the location of the neutral point), and Sees, 5.3 and 5.10.

THE METHOD OF NORMAL MODES

When the separation in frequency between the elastic degrees of freedom and therigid-body motions is not large, then significant inertial coupling can occur betweenthe two. In that case a dynamic analysis is required, which takes account of the timedependence of the elastic motions.

The method that is described here for accomplishing this purpose is based uponthe representation of the deformation of the elastic vehicle in terms of its normalmodes of free vibration. Imagine that the vehicle is at rest under the action of no ex-ternal forces, aerodynamic, gravity, or other, and that a frame of reference with originat the mass center, but otherwise arbitrary, is attached to it. The position of mass ele-ment Bm is then (xo, Yo, zo). Now let the structure be deformed by a self-equilibratingset of external forces and couples, so that it takes a new form, stationary with respectto the coordinate system. Upon instantaneous release of this force system, a free vi-bration ensues, that is, one in which external forces play no part, and in which the po-sition of 5m at time t is (x, y, z). Since there is zero net force, and zero net moment,the linear and angular momenta of the elastic motion must vanish, whatever the ini-tial distorted shape. In particular this is true for each and every undamped normalmode of free vibration. Any small arbitrary elastic motion of the vehicle can, there-fore, relative to the chosen axes (transients as well as steady oscillations), be repre-sented by a superposition of free undamped normal modes as follows:

ec

x'(t) =I fn(xo, Yo, Zo)€n(t)1

00

y'(t) = I gn(XO' Yo, ZO)€n(t)1

(4.12,1)

00

z' (t) =I hn(xo, Yo, ZO)€n(t)1

Page 132: Stability Control EtkinReid

122 Chapter 4. General Equations of Unsteady Motion

where (x', y', z') are the elastic displacements, (x - xo) etc., (f n' gn' hn) are the modeshape functions;' and En(t) are the generalized coordinates giving the magnitudes ofthe modal displacements.

We have specified idealized undamped modes, as opposed to the true modes of areal physical structure with internal and external damping, because the latter may notbe "simple" modes with fixed nodes, describable by a single set of three functions.More generally they each consist of a superposition of two "submodes" 90° out ofphase. Because of this, the equations of motion for the elastic degrees of freedom ofthe real structure are not perfectly uncoupled from one another, but contain intercou-pling damping terms that would usually be negligible in practical applications.

The use of the free undamped normal modes is seen to ensure that the linear andangular momenta of the distortional motion vanishes. Consequently the elastic mo-tions have no inertial coupling with the rigid-body motions except through the mo-ments and products of inertia. However, it can be shown that this coupling is second-order and negligible in the small-perturbation theory. The determination of the shapesand frequencies Wn of the normal modes is a major task, and the methods for findingthem are beyond the scope of this text. For treatises on this subject the reader shouldrefer to (Bisplinghoff et al., 1955; Fung, 1955). As indicated in (4.12,1), there are ac-tually an infinite number of normal modes of vehicle structures. In practice, ofcourse, only a finite number N of those at the low-frequency end of the set need beretained, and the summations in (4.12,1) are approximated by finite series of N terms.Some judgment and experience is needed to decide just how many modes are neededin any application, but a general rule that is helpful is to discard those whose frequen-cies are substantially higher than the significant ones present in the spectral represen-tation of inputs arising from control action or atmospheric turbulence.

MODIFICATION OF THE RIGID BODY EQUATIONS

Although the inertia terms of the previous equations, for example, (4.9,18) and(4.9,19), remain unchanged to first order by the presence of elastic motions, the elas-tic and rigid-body motions are not nevertheless entirely uncoupled.

The deformations of the structure in general cause perturbations in the aerody-namic forces and moments. These may be introduced into the linearized equations ofmotion by the addition of appropriate derivatives to the expressions for the aerody-namic forces given by (4.9,17). For example, the added terms in the pitching momentassociated with the nth elastic degree of freedom would be

(4.12,2)

Similar expressions appear for each of the added degrees of freedom, and in each ofthe aerodynamic force and moment equations. An example of the elastic stability de-rivatives is given in Sec. 5.10. Alternatively, the aerodynamic forces may be formu-lated in the form of transfer functions.

THE ADDITIONAL EQUATIONS OF MOTIONThe additional equations are most conveniently found by using Newton's laws as ex-pressed by Lagrange's equations of motion (Synge and Griffith, 1942) with the En as

'The eigenfunctions of the linear vibration problem.

Page 133: Stability Control EtkinReid

4.12 Elastic Degrees of Freedom 123

generalized coordinates. The appropriate form of Lagrange's equation for this appli-cation is

(4.12,3)

where T is the kinetic energy of the elastic motion relative to FB, U is the elasticstrain energy, and '!:Fn is the generalized external force. Since the coordinates are mea-sured in the frame FB' which is non-Newtonian by virtue of its general motion, an ap-propriate modification must be made to the external force field acting on the systemwhen calculating the generalized force. This consists of adding to each element ofmass am an inertial body force equal to - a I am where a' is that part of the total ac-celeration of am that arises from the acceleration and rotation of FB (see AppendixA.6).

Since normal modes have been chosen as the degrees of freedom, then the indi-vidual equations of motion are independent of one another insofar as elastic and iner-tia forces are concerned (this is a property of the normal modes), although the equa-tions will be coupled through the aerodynamic contributions to the '!:F's. The lack ofelastic and inertia coupling permits the left-hand side of (4.12,3) to be evaluated byconsidering only a single elastic degree of freedom to be excited. Let its generalizedcoordinate be En' The kinetic energy is given by

where the integration is over all elements of mass of the body. From (4.12,1) this be-comes (with only En excited)

The integral is the generalized inertia in the nth mode, and is denoted by

(4.12,4)

so that

(4.12,5)

The first term of (4.12,3) is therefore InEn, and the second term is zero.The strain-energy term is conveniently evaluated in terms of the natural fre-

quency of the nth mode by applying Rayleigh's method. This uses the fact that, whenthe system vibrates in an undamped normal mode, the maximum strain energy occurswhen all elements are simultaneously at the extreme position, and the kinetic energyis zero. This maximum strain energy must be equal to the maximum kinetic energythat occurs when all elements pass simultaneously through their equilibrium position,where the strain energy is zero. Hence, if En = a sin (lV, then the maximum kineticenergy is, from (4.12,5)

Page 134: Stability Control EtkinReid

124 Chapter 4. General Equations of Unsteady Motion

Since the stress-strain relation is assumed to be linear, the strain energy" is a qua-dratic function of E"n; that is, U = ik~. Hence

Uma:x =: ika2 = Tma:x = iIn~a2

It follows that k = In(J)~, and that

u = iIn(J)~~

and hence aUlaE"n = In(J)~E"n' The left side of (4.12,3) is therefore as follows:

(4.12,6)

When structural damping is present, this simple form of uncoupled equation isnot exact but the changes in frequency and mode shape for small damping are notlarge. Hence damping can be allowed for approximately by adding a damping term to(4.12,6), that is,

(4.12,7)

without changing (J)n or the mode-shape functions. The value of , is ordinarily lessthan 0.1, and usually must be found by an experimental measurement on the actualstructure.

EVALUATION OF ~n

The generalized force is calculated from the work done during a virtual displacement,

aw'?Ji'n = ~

oe;(4.12,8)

where W is the work done by all the external forces, including the inertia forces asso-ciated with nonuniform motion of the frame of reference. The inertia force field isgiven by

dfj = -(a') dm (4.12,9)

where the components of the r.h.s. in Fa are given by (A.6,8) without the terms (x,y,z). The work done by these forces in a virtual displacement of the structure is

where the integration is over the whole body. Introducing (4.12,1) this becomes

whence (4.12,10)

When the inertia-force expressions are linearized to small disturbances, and substi-tuted into (4.12,10), all the remaining first-order terms contain integrals of the fol-lowing types:

8Por example, in a spring of stiffness k and stretch x, the strain energy is U = !A;x2.

Page 135: Stability Control EtkinReid

4.12 Elastic Degrees of Freedom 125

The first of these is zero because the origin is the mass center, and the second is zerobecause the angular momentum associated with the elastic mode vanishes. The netresult is that 'i!Fni = O.This result simply verifies what was stated above; that is, thereis no inertial coupling between the elastic and rigid-body degrees of freedom.

The remaining contribution to 'i!Fn is that of the aerodynamic forces. Let the localnormal-pressure perturbation at an element dS of the airplane's surface be p(xo, Yo,Zo), and let the local outward normal be n(nx' ny, nz). Then the work done by the aero-dynamic forces in a virtual displacement is

SWa = - f pn . (r - ro)dS

where the integral is over the whole surface of the airplane, and (r - ro) is the vectordisplacement at dS. It is given by

00

r - ro =I (ifn + jgn + khn) SEnn=l

hence

8Wa = - i: 8En f p(nxfn + n~n + nzhn) dSn=l

and

(4.12,11)

Each of the variables inside the integral is a function of (xo, Yo, Zo), i.e., of position onthe surface, and moreover, p is in the most general case a function of all the general-ized coordinates, of their derivatives, and of the control-surface angles. The result isthat 'i!Fn is a linear function of all these variables, which may be expressed in terms ofa set of generalized aerodynamic derivatives (or alternatively aerodynamic transferfunctions), namely,

'i!F = A Au + A ·it + ... + A_n + ... + A 8 + ...n nu nu npr nl}r r

00 00 00

+ I anmEm + I bnmEm + I CnmEmm=l m=l m=l

(4.12,12)

In application, only the important derivatives would be retained in any given case.The values of the derivatives kept would be computed by application of (4.12,11). Anexample of this computation is given in Sec. 5.10.

RESUME

The effects of structural dynamics on the stability and control equations can be incor-porated by adding structural degrees of freedom based on free normal modes. For anexact representation, an infinite number of such modes are required; however, in

Page 136: Stability Control EtkinReid

126 Clutpter 4. General Equations of Unsteady Motion

practice only a few of the lowest modes need be employed. The six rigid-body equa-tions are altered only to the extent of additional aerodynamic terms of the type givenin (4.12,2). One additional equation is required for each elastic degree of freedom(4.12,7). The generalized forces appearing in the added equations contain only aero-dynamic contributions, which are computed from (4.12,11) and expressed as in(4.12,12).

4.13 Exercises4.1 Prove that the angular momentum vector h of an airplane is the same whether or not

the wind vector W is zero.

4.2 Carry out the expansion of fwr to derive the result given in (4.3,6).

4.3 Carry out the multiplication of the three rotation matrices indicated in (4.4,2) to ob-tain the result given in (4.4,3).

4.4 Derive (4.5,11) for the moments and product of inertia in frame Fs when they aregiven for principal axes.

4.5 Prove that when an airplane has spinning rotors, such as jet engines, the angular mo-mentum is given by (4.6,1). Derive the additional terms in the moment equationsgiven in (4.6,2).

4.6 Two airplanes are geometrically similar and have similar mass distributions. AirplaneA has a span of 100 ft (30.48 m) and a weight of 100,000 lb (445,000 N). B has 150-ft(47.72 m) span and weighs 225,000 lb (1,001,250 N). Both fly at speeds low enoughto neglect Mach number effects, and high enough to neglect Reynolds number ef-fects.

When flying at 400 knots and 20,000-ft (6,096-m) altitude, airplane B has a spi-ral divergence (a lateral instability) that has a characteristic time of 20 seconds.

(a) At what speed and altitude will A be dynamically similar to B?(b) What will be the characteristic time of the spiral divergence of A at that speed

and altitude?(c) What is the ratio of the CL values for the two flight conditions?

4.7 Substitute the linear expressions for til, and AM into the right side of (4.9,7c) and(4.9,8b) and solve the resulting equations to get the second and third components of(4.9,18).

4.8 Derive the expressions given in Tables 4.4 and 4.5 for the derivatives Xq, Mw, Yp, Nv'

4.9 Let [PE qE rE]Tbe the angular velocity of an airplane in frame FE' Find a set of equa-tions that produces these components, given the body axis components [p q rf.

4.10 A hovercraft in ground effect is acted on by the following aerodynamic forces, ex-pressed as body frame components:

X=Y=OZ= -mg + ZzZE

L = L."cP; M = M lJ; N = 0

Page 137: Stability Control EtkinReid

4.14 Additional Symbols Introduced in Chapter 4 127

The body axes are principal axes, and the engine/rotor angular momentum is

h~= [0 0 Hf

Derive a set of small-disturbance equations of motion.

Suppose that the wind vector W is not zero but instead is given by

WB = rUg vg wgf

where W is assumed to be "small" and a known function of time. Derive the addi-tional terms that would have to be added to the small-disturbance equations of mo-tion (4.9,18) and (4.9,19).

4.12* An aircraft is performing a rolling pullup. At the instant of observation, the vehicle isat the bottom of a vertical circle of 2000 ft (610 m) radius moving at a constant speedof 500 fps (152 mls) with wings horizontal. (See Fig. 3.1). At the same time the rollrate is constant at p = 90° S-1. Given that

4.11

Iy - I, = 300 slug fe (406 kg m2) and I,= 500 slug ft2 (677 kg m2

)

determine the moments required at this time to perform this maneuver. Assume thatthe axes are principal axes, with ex horizontal. (You may assume constant Euler an-gle rates and l/J = 0.)

Cwot; t; t,I~,I~Iyz

Ixz

I~Ixy

sGh

h'

(hx' hy, hz)

(h~, »; h~)

(L,M,N)

(p, q, r)

4.14 Additional Symbols Introduced in Chapter 4

mg/ipuo2S

moments of inertia about (x, y, z) axessee (4.9,19)product of inertia f yz dmproduct of inertia f xz dmsee (4.9,19)product of inertia f .xy dmresultant external force vectorgeneralized force in Lagrange's equationresultant external moment vector, about the mass centerangular momentum vector of the airplane with respect to its masscenterangular momentum vector of spinning rotors with respect to ro-tor mass centerscalar components of h in FB

scalar components ofh' in FB

scalar components of G in FB

scalar components of Cd, radians/sec in FB

*Problem courtesy of Prof. F. H. Lutze, Virginia Polytechnic Institute

Page 138: Stability Control EtkinReid

128 Chapter 4. General Bquation« of Unsteady Motion

(u, v, w) scalar components of V in FB

airspeed vector of airplane mass centercomponents of resultant aerodynamic force acting on the air-plane, inFB

coordinates of airplane mass center relative to fixed axes (seeFig. 4.2)

En generalized coordinate of the nth elastic mode5e, 5,.,5a angles of elevator, rudder, and aileron5p propulsion controlCJJ angular velocity vector of the airplane(t/J, (), </J) Euler angles, radians (see Sec. 4.4)

See also Tables 4.1 to 4.3.See also Sees. 2.11 and 3.15.

V(X, Y, Z)

Page 139: Stability Control EtkinReid

CHAPTER 5

The Stability Derivatives

5.1 General RemarksWe saw in Chap. 4 how the aerodynamic actions on the airplane can be representedapproximately by means of stability derivatives (or more exactly by aerodynamictransfer functions). Indeed, all the aerodynamics involved in airplane dynamics isconcentrated in this section of the subject: i.e., in the determination of these deriva-tives (or transfer functions). Each of the stability derivatives contained in the equa-tions of motion is discussed in the following sections. Wherever possible, formulasfor them are given in terms of the more elementary parameters used in static stabilityand performance. Where this is not feasible, it is shown in a qualitative way how theparticular force or moment is related to the relevant perturbation quantity. No data forestimation are contained in this chapter; these are all in Appendix B.

EXPRESSIONS FOR c,AND c,For convenience, we shall want the derivatives of C; and C, expressed in terms of lift,drag, and thrust coefficients. The relevant forces are shown in Fig. 5.1. As shown, thethrust line does not necessarily lie on the x axis. However, the angle between them isgenerally small, and we shall assume it be zero. With this assumption, and for smallax, we get'

C; = CT + CLax - CD

c,= -(CL + CDax)(5.1,1)

where CT is the coefficient of thrust, Tlip V2s.

•• 5.2 TheaDerivatives (CXa' Cza' Cm)The a derivatives describe the changes that take place in the forces and momentswhen the angle of attack of the airplane is increased. They are normally an increasein the lift, an increase in the drag, and a negative pitching moment. The contents ofChap. 2 are relevant to these derivatives.

'Since X and Z are the aerodynamic forces acting on the airplane, there are no weight components in(5.1,1).

129

Page 140: Stability Control EtkinReid

130 Chapter 5. The Stability Derivatives

L

• wFigure 5.1 Forces in symmetric flight.

THE DERIVATIVE c;By definition, Cx" == (aCxlda)o, where the subscript zero indicates that the derivativeis evaluated when the disturbance quantities are zero. From (5.1,1)

sc, dCT dCL dCnda == da + CL + ax da - da

We may assume that the thrust coefficient is sensibly independent of ax so thatdCTlda == 0, and hence

..

C == (dCx) == C _ (dCn)Xa :::Ia Lo dau 0 0

(5.2,1)

where the subscript zero again indicates the reference flight condition, in which, withstability axes, ax == O.When the drag is given by a parabolic polar in the form Cn ==Cnmin + Ci!TrAe, then

(dCn) == 2CLo CLda ° TrAe a

(5.2,2)

THE DERIVATIVE CZa

Bydefinition, CZa == (dC/da)o. From (5.1,1) we get

dCz ( acn)da == - CLa + Cn + ax da

Therefore

CZa == -(CLa + Cno) (5.2,3)

Cno will frequently be negligible compared to CLa' and consequently CZa~ -CLa'

THE DERIVATIVE c.;Cma is the static stability derivative, which was treated at some length in Chap. 2. It isconveniently expressed in terms of the stick-fixed neutral point (2.3,25):

Page 141: Stability Control EtkinReid

5.3 The u Derivatives tc.; C,.' Cm) 131

Cmu = a(h - hn)

For airplanes with positive pitch stiffness, h < hn, and Cme<is negative.

(5.2,4)

5.3 The u Derivatives tc., c., Cm)The u derivatives give the effect on the forces and moments of an increase in the for-ward speed, while the angle of attack, the elevator angle, and the throttle position re-main fixed. If the coefficients of lift and drag did not change, then this would implyan increase in these forces in accordance with the speed-squared law, i.e.,

Force or moment (uo + AU)2-------- = 2 =1= I + 2AuInitial force or moment Uo

Since the pitching moment is initially zero, then, so long as Cm does not change withu, it will remain zero. The situation is actually more complicated than this, for thenondimensional coefficients are in general functions of Mach number and Reynoldsnumber, both of which increase with increasing u. The variation with Reynolds num-ber is usually neglected, but the effect of Mach number must be included.

The thrust effect shows up in two different ways. One stems simply from the de-rivative of thrust with speed, which depends on the type of propulsive system-jet,propeller, and so forth. The other, related mainly to propeller configurations, derivesfrom the propulsion/airframe interaction, for example, the propeller slipstream im-pinging on the wing. This is an important effect, and for some STOL airplanes, maybe dominant at low speeds.

Finally, the increased loading on the airframe due to the speed increase may in-duce significant structural distortion. This is a static aeroelastic effect. For example,the tail lift coefficient may be influenced appreciably by the loading (see Sec. 3.5).An appropriate variable to use for aeroelastic effects is the dynamic pressure Pd =~pV2.

In order to formally include each of these three major effects, compressibility,aeroelasticity, and propulsive, even though they would rarely all be present at thesame time, each of the coefficients Cx, CZ' Cm is assumed to be a function of M, Pd'and CT as well as angle of attack.

We then have

C = (acx) = (acx aM) + (acx aPd) + (acx aCT)Xu au 0 aM au 0 apd au 0 aCT au 0

and similarly for Czu and Cmu'(5.3,1)

CALCULATION OF ews«The Mach number is M = VIa, where a is the speed of sound, so

aM aM Uo sv avau = Uo au = -; a; = Mo au

But

Page 142: Stability Control EtkinReid

132 Chapter 5. The Stability Derivatives

so (:t = 1

(O~) = Meou 0

(5.3,2)and

CALCULATION OF iJpdf{JU

so

and (OPd) == ( (JPd) = 2~A Uo ~ PUoou 0 ou 0

The thrust derivative is defined in a manner consistent with Table 4.2 to be

(5.3,3)

(JCTC~ =--;:;-

• aU(5.3,4)

It follows that Cx• is given by

(5.3,5)

THE DERIVATIVE CT.

Since

(JCT aT/au 2T svthen a;- == ipv2s - ipV3S auIn the reference flight condition V = Uoand av/ou = 1, so

_ (OCT) _ (dCT) _ «(JT/ou)o _CT. - ~ - Uo ~ - 1 •• ~ 2CToau 0 ou 0 lPH(JU

For unpowered gliding flight, T = 0 and

CT. = 0 (5.3,7)

For constant-thrust propulsion, which is an approximation for jet aircraft in cruisingflight, aT/(Ju == 0, and

(5.3,6)

CT. = -2CTo (5.3,8)

For constant-power propulsion, which is an approximation for constant-speed pro-pellers in cruising flight, TV is constant, so that

(oTlou)o = -Toluoand CT. = -3CTo (5.3,9)

The values of CToin the preceding expressions can be related to the reference lift and

Page 143: Stability Control EtkinReid

5.3 The u Derivatives (Cxu' CZu' Cm) 133

drag coefficients [see Fig. 5.1 and (5.1,1)]. Note that T, V, x are assumed to be colin-ear, that is,

(5.3,10)

THE DERIVATIVE c;From (5.1,1) we have

(dCX) = (dCT) _ (dCD)

dM 0 dM 0 dM 0

(dCx) (dCT) (dCD)dPd 0 = dPd 0 - dPd 0

(dCx) (dCD)dCT 0 = 1 - dCT 0

Since the direct aeroelastic effect on thrust is likely to be negligible, we neglectdCTldPd' and then (5.3,5) gives

c; = Mo(~~ - ~; t -pu~ (~~;)o+ CTu (1 - ~~: t (5.3,11)

When a powered wind-tunnel model is tested, it is common practice to measure thenet axial force coefficient C; and not its component parts CT and CD' In that case, thetest data can provide the Cxu derivative directly.

THE DERIVATIVE CZu

From (5.1,1) we have that

(dCZ) = _(dCL)

dM 0 dM 0

(dCz) (dCL)dPd 0 = - dPd 0

( dCz) (dCL)dCT 0 = - dCT 0

so that CZu = -Mo (~~ t -pu~ (~~; t -CTu (~~;t (5.3,12)

The derivative Mo(dCddMo) tends to be small except at transonic speeds. Theoreticalvalues are easily calculated for high-aspect-ratio swept wings. At subsonic speeds,the Prandtl-Glauert rule combined with simple sweep theory (Kuethe and Chow,1976) gives the lift coefficient for two-dimensional flow as

aj(XC - McosA< 1

L - y'1 - M2 cos2 A

where a, is the lift-curve slope in incompressible flow and A is the sweepback angleof the ~chord line. Upon differentiation with respect to M, we get

Page 144: Stability Control EtkinReid

134 Chapter 5. The Stability Derivatives

and hence

(OCL) ~ cos2 A C

Mo oM 0 = 1 - ~ cos'' A La (5.3,13)

In level flight with the lift equal to the weight, M~CLo is constant, and henceMo(oCJoM)o is proportional to 1/(1 - ~ cos? A). At supersonic speeds, the two-di-mensionallift is given by Kuethe and Chow (1976)

4acos AC - ~::::::;;=====;:;:=:==

L - YM2 cos" A-I

After differentiation with respect to M, we get exactly the same result as for subsonicspeeds. That is (5.3,13) applies over the whole Mach-number range, except of coursenear M = 1 where the cited airfoil theories do not apply. Low-aspect-ratio wings areless sensitive to changes in M.

THE DERIVATIVE em"From (5.3,5) and (5.3,6) Cm" is given as

_ (OCm) 2 (OCm) (OCm)Cm• - M aM 0 + PUo apd 0 + CT. aCT 0 (5.3,14)

Values of aC,joM can be found from wind-tunnel tests on a rigid model. They arelargest at transonic speeds and are strongly dependent on the wing planform. Themain factor that contributes to this derivative is the backward shift of the wing centerof pressure that occurs in the transonic range. On two-dimensional symmetricalwings, for example, the center of pressure moves from approximately O.25c to ap-proximately O.50c as the Mach number increases from subsonic to supersonic values.Thus an increase in M in this range produces a diving-moment increment; that is, Cm•

is negative. For wings of very low aspect ratio, the center of pressure movement ismuch less, and the values of Cm• are correspondingly smaller.

To find OCiapd requires either an aeroelastic analysis or tests on a flexiblemodel. As an example of this phenomenon, let us consider an airplane with a tail anda flexible fuselage.' We found in Sec. 3.5 that the tail lift coefficient is given by

atCL,= ---- (awb - E - it)

1 + katPdSt

The pitching moment contributed by the tail is (2.2,9)

em, = -VeCL,

(5.3,15)

Hence

(5.3,16)

2It is not meant to imply that fuselage bending is the only important aeroelastic contribution to em.'Distortion of the wing and tail may also be important.

Page 145: Stability Control EtkinReid

5.4 The q Derivatives (CZq' Cmq) 135

When (5.3,15) is differentiated with respect to Pd and simplified, and the resulting ex-pression is substituted into (5.3,16), we obtain the result

(OCm) katSt0Pd tail = -Cmt 1+ katPdSt

The corresponding contribution to Cmu is [see (5.3,14)]

2PdokarSt(CmJtail = -Cmt 1 + ka S

tPdo t

All the factors in this expression are positive, except for Cmt' which may be of eithersign. The contribution of the tail to Cmu may therefore be either positive or negative.The tail pitching moment is usually positive at high speeds and negative at lowspeeds. Therefore its contribution to Cmu is usually negative at high speeds and posi-tive at low speeds. Since the dynamic pressure occurs as a multiplying factor in(5.3,18), then the aeroelastic effect on Cmugoes up with speed and down with altitude.

(5.3,17)

(5.3,18)

5.4 The q Derivatives (CZq' Cm)These derivatives represent the aerodynamic effects that accompany rotation of theairplane about a spanwise axis through the CG while ax remains zero. An example ofthis kind of motion was treated in Sec. 3.1 (i.e., the steady pull-up). Figure S.2bshows the general case in which the flight path is arbitrary. This should be contrastedwith the situation illustrated in Fig. S.2a, where q = 0 while ax is changing.

la)

(hI

Figure 5.2 (a) Motion with zero q, but varying ax. (b) Motion with zero ax but varying q.

Page 146: Stability Control EtkinReid

136 Chapter 5. The Stability Derivatives

Both the wing and the tail are affected by the rotation, although, when the air-plane has a tail, the wing contribution to CZqand Cmq is often negligible in compari-son with that of the tail. In such cases it is common practice to increase the tail effectby an arbitrary amount, of the order of 10%, to allow for the wing and body.

CONTRIBUTIONS OF A TAIL

As illustrated in Fig. 5.3, the main effect of q on the tail is to increase its angle of at-tack by (ql/uo) radians, where Uo is the flight speed. It is this change in at that ac-counts for the changed forces on the tail. The assumption is implicit in the followingderivations that the instantaneous forces on the tail correspond to its instantaneousangle of attack; i.e., no account is taken of the fact that it takes a finite time for thetail lift to build up to its steady-state value following a sudden change in q. (Amethod of including this refinement has been given by Tobak, 1954.) The derivativesobtained are therefore quasistatic.

CZqof the TailBy definition, CZq = (ac/aq)o = (2uo!c)(ac/aq)o, and, from (5.1,1), (ac/aq)o

= -(aCdaq)o' The change intaillift coefficient caused by the rotation q is

q1t!:i.CL, = a,l1at = at-Uo

(5.4,1)

and the corresponding change in airplane lift coefficient is

sc = St!:i.C = St a q1tL S Lt S tuo

v

v

Zero lift line

Figure 5.3 Effect of pitch velocity on tail angle of attack.

Page 147: Stability Control EtkinReid

5.4 The q Derivatives (C:<q'Cmt) 137

Therefore

and

(5.4,2)

em of the Tallq

The increment in pitching moment that corresponds to ACL, is [see (2.2,9)]

qltACm = -VHACL, = -atVH-

Uo

Hence

aACm It-- = -aVH-aq t Uo

and

2uo ( sse; ) It(Cm,)tail = --- --a- = -2at VH-=-

C q 0 c(5.4,3)

CONTRIBUTIONS OF A WINGAs previously remarked, on airplanes with tails the wing contributions to the q deriv-atives are frequently negligible. However, if the wing is highly swept or of low aspectratio, it may have significant values of CZq and Cmq; and of course, on tailless air-planes, the wing supplies the major contribution. The q derivatives of wings alone aretherefore of great engineering importance.

Unfortunately, no simple formulas can be given, because of the complicated de-pendence on the wing planform and the Mach number. However, the following dis-cussion of the physical aspects of the flow indicates how linearized wing theory canbe applied to the problem. Consider a plane lifting surface, at zero ax, with forwardspeed Uo and angular velocity q about a spanwise axis (see Fig. 5.4). Each point inthe wing has a velocity component, relative to the resting atmosphere, of qx normalto the surface. This velocity distribution is shown in the figure for the central and tipchords. Now there is an equivalent cambered wing that would have the identical dis-tribution of velocities normal to its surface when in rectilinear translation at speed Uo'This is illustrated in Fig. 5.5a. The cross section of the curved surface S is shown in(b). The normal velocity distribution will be the same as in Fig. 5.4 if

az oz qUo - = qx or - = - xax ax Uo

Hence

1 qz=--x22 Uo

(5.4,5)

Page 148: Stability Control EtkinReid

138 Chapter 5. The Stability Derivatives

///

II

Figure 5.4 Wing velocity distribution due to pitching.

and the cross section of S is a parabolic arc. In linearized wing theory, both subsonicand supersonic, the boundary condition is the same for the original plane wing withrotation q and the equivalent curved wing in rectilinear flight. The problem of findingthe q derivatives then is reduced to that of finding the pressure distribution over theequivalent cambered wing. Because of the form of (5.4,5), the pressures are propor-tional to qluo. From the pressure distribution, CZq and Cmq can be calculated. The de-rivatives can in principle also be found by experiment, by testing a model of theequivalent wing.

The values obtained by this approach are quasistatic; i.e., they are steady-statevalues corresponding to ax = 0 and a small constant value of q. This implies that theflight path is a circle (as in Fig. 3.1), and hence that the vortex wake is not rectilinear.Now both the linearized theory and the wind-tunnel measurement apply to a straightwake, and to this extent are approximate. Since the values of the derivatives obtainedare in the end applied to arbitrary flight paths, as in Fig. 5.2b, there is little point incorrecting them for the curvature of the wake.

The error involved in the application of the quasistatic derivatives to unsteadyflight is not as great as might be expected. It has been shown that, when the flightpath is a sine wave, the quasistatic derivatives apply so long as the reduced frequencyis small, that is,

wek = - ~ 1 (5.4,6)

2uowhere lIJ is the circular frequency of the pitching oscillation. If 1is the wavelength ofthe flight path, then

Ck= 'TT-

I

so that the condition k ~ 1 implies that the wavelength must be long compared to thechord, for example, I > 60c for k < 0.05.

Page 149: Stability Control EtkinReid

5.4 The q Derivatives (CZq' Cm,) 139

"0

s(al

II" o

"0

(bl

Figure 5.5 The equivalent cambered wing.

DEPENDENCE ON h

Because the axis of rotation in Fig. 5.5, passes through the CG, the results obtainedare dependent on h. The nature of this variation is found as follows. Let the axis ofrotation be atA in Fig. 5.6, and let the associated lift and moment be

(5.4,7)

Now let the axis of rotation be moved to B, with the change in normal velocity distri-bution shown on the figure. Since the two normal velocity distributions differ by aconstant, (the upward translation qc Ah)the difference between the two pressure dis-tributions is that associated with a flat plate at angle of attack

qca= --Ah

Uo(5.4,8)

This angle of attack introduces a lift increment acting at the wing mean aerodynamiccenter of amount

(5.4,9)

Page 150: Stability Control EtkinReid

140 Chapter 5. The Stability Derivatives

••

~Czq

-+---..,.,.....,~••.••:;..-.---h~ ....l __

(b)

Figure 5.6 Effect of CO location on CZq' Cmq.

aCL I:i.CL

oh = I:i.h = - 2CLaq

02CL 0ohoq = oh CLq = -2CLa

We see that CL is linear in h, and can therefore be expressed asq

that is,

and (5.4,10)

CLq = -2CLa(h - ho) (5.4,11)

where ho is the CG location at which C""is zero. By virtue of (5.1,1) we get

CZq= -CLq = 2CLa(h - ho) (5.4,12)

The pitching moment about the CG is

c; = Cm•c + CL(h - hn) (5.4,13)

so thatoCmoc

Cm = ~ + CT(h - h; )q dq -. W

oCmac= -- - 2C (h - h )(h - h )oq La 0 nw

(5.4,14)

Equation (5.4,14) shows that Cmq is quadratic in h. We can write it without loss ofgenerality as

Cmq = Cmq - 2CLJh - Iii (5.4,15)

where Cmq is the maximum (least negative) v~ue of Cmq and Ii is the CG locationwhere it occurs (see Fig. 5.6b). The value of h is found by differentiating (5.4,14).This yields

(5.4,16)

Page 151: Stability Control EtkinReid

5.5 The aDerivatives (CL••, Cm..) 141

The linear theory of two-dimensional thin wings gives for supersonic flow:

ho=Ji=~_ 2Cmq = - 3YM2 - 1 (5.4,17)

and for subsonic flow:

ho =!h--1-2

C =0mq

(5.4,18)

PITCH DAMPING OF PROPULSIVE JETSWhen gases flow at high speed inside jet or rocket engines at the same time as the ve-hicle is rotating in pitch or yaw, they react against the walls of the ducts WItha forceperpendicular to their velocity vector (the Coriolis force). This reaction can result in apitching moment proportional to q, that is, in a contribution to Cmq, (and similarly toCn).An analysis of this effect is given in Sec. 7.9 ofEtkin (1972).

For jet airplanes in cruising flight this contribution to Cmq is usually negligible.Only at high values of CT> and when the Cmq of the rest of the airplane is small,would it be significant. On the other hand, a rocket booster at lift-off, when the speedis low, has practically zero external aerodynamic damping and the jet damping be-comes very important.

5.5 The aDerivatives (CLM Cm)

The Ct derivatives owe their existence to the fact that the pressure distribution on awing or tail does not adjust itself instantaneously to its equilibrium value when theangle of attack is suddenly changed. The calculation of this effect, or its measure-ment, involves unsteady flow. In this respect, the Ct derivatives are very different fromthose discussed previously, which can all be determined on the basis of steady-stateaerodynamics.

CONTRIBUTIONS OFA WING

•Consider a wing in horizontal flight at zero a. Let it be subjected to a downward im-pulse, so that it suddenly acquires a constant downward velocity component. Then, asshown in Fig. 5.7, its angle of attack undergoes a step increase. The lift then respondsin a transient manner (the indicial response) the form of which depends on whetherM is greater or less than 1. In subsonic flight, the vortices which the wing leaves be-hind it can influence it at all future times, so that the steady state is approached onlyasymptotically. In supersonic flight, the upstream traveling disturbances move moreslowly than the wing, so that it outstrips the disturbance field of the initial impulse ina finite time tt- From that time on the lift remains constant.

In order to find the lift associated with Ct, let us consider the motion of an airfoilwith a small constant Ct, but with q = O. The motion, and the angle of attack, are

Page 152: Stability Control EtkinReid

142 Chapter 5. The Stability Derivatives

f

1:1------~O"'------------.sore

Asymptote=:;:;.-.._----c.----M.O

-------I~---------------.eo

OLt

~-~>o-----~o • f

OLtJ( Steady-state value

~~ll______ ~L~ . •~o t1

Figure 5.7 Lift response to step change in a. (After Tobak, NACA Rept. 1188.)

shown in Fig. 5.8. The method used follows that introduced by Tobak (1954). We as-sume that the differential equation which relates CL(i) with act) is linear. Hence themethod of superposition (the convolution integral) may be used to derive the re-sponse to a linear aCt). Let the response to a unit step be A(t). Then the lift coefficientat time t is (see Appendix A.3).

CL(i) = f=o A(t - 'T)a( 'T) dr

•••

Since a( 'T) = constant, then

(5.5,1)

Page 153: Stability Control EtkinReid

5.5 The aDerivatives (CL••, Cm) 143

o

(0)

I'---------~t

"--------- •. t

(c)

Figure 5.8 Lift associated with a.

The ultimate CL response to a unit-step a input is CLa• Let the lift defect be f(t):that is,

A(f) = CLa - f(f)

Then (5.5,1) becomes

CL(f) = aCL} - a f=o f(f - T) dr

= CLaa - Sa (5.5,2)

where S(f) = f;=o f(f - T) dt: The term S a is shown on Fig. 5.8. Now, if the idea ofrepresenting the lift by means of aerodynamic derivatives is to be valid, we must beable to write, for the motion in question,

(5.5,3)

Page 154: Stability Control EtkinReid

144 Chapter 5. The Stability Derivatives

1m

Figure 5.9

where CLa and CL", are constants. Comparing (5.5,2) and (5.5,3), we find that CLi>.=-S(f), a function of time. Hence, during the initial part of the motion, the derivativeconcept is invalid. However, for all finite wings," the area S(f) converges to a finitevalue as f increases indefinitely. In fact, for supersonic wings, S reaches its limitingvalue in a finite time, as is evident from Fig. 5.7. Thus (5.5,3) is valid," with constantCLa' for values of f greater than a certain minimum. This minimum is not large, beingthe time required for the wing to travel a few chord lengths. In the time range whereS is constant, or differs only infinitesimally from its asymptotic value, the CL(f) curveof Fig. 5.8c is parallel to CLa a. A similar situation exists with respect to Cm'

We see from Fig. 5.8 that CL"" which is the lim - S(f), can be positive for M = 0and negative for larger values of M. Hoc

There is a second useful approach to the it derivatives, and that is via considera-tion of oscillating wings. This method has been widely used experimentally, and ex-tensive treatments of wings in oscillatory motion are available in the literature," pri-marily in relation to flutter problems. Because of the time lag previously noted, theamplitude and phase of the oscillatory lift will be different from the quasisteady val-ues. Let us represent the periodic angle of attack and lift coefficient by the complexnumbers

and (5.5,4)

where ao is the amplitude (real) of a, and CLo is a complex number such that IcLoI isthe amplitude of the CL response, and arg C1.0 is its phase angle. The relation betweenCLoand ao appropriate to the low frequencies characteristic of dynamic stability is il-lustrated in Fig. 5.9. In terms of these vectors, we may derive the value of CL'" as fol-lows. The it vector is

Thus CL may be expressed as

CL = R[CLo]eiwt + iI[CLo]eiwt

3Por two-dimensional incompressible flow, the area S(c) diverges as t ~ 00. That is, the derivativeconcept is definitely not applicable to that case.

"Bxactly for supersonic wings, and approximately for subsonic wings.sSee bibliography.

••

Page 155: Stability Control EtkinReid

5.5 The aDerivatives (CL;,,'Cm) 145

CL2

I CGex ... 0

M~

tcV \

c

Figure 5.10 Lift on oscillating two-dimensional airfoil.

Hence (5.5,5)

or, if the amplitude ao is unity, CLix = /[CLo]/k, where k is the reduced frequencywel2uo·

To assist in forming a physical picture of the behavior of a wing under these con-ditions, we give here the results for a two-dimensional," airfoil in incompressibleflow. The motion of the airfoil is a plunging oscillation; that is, it is like that shown inFig. 5.2a, except that the flight path is a sine wave. The instantaneous lift on the air-foil is given in two parts (see Fig. 5.10):

CL = CLI + CL2

CLI = 21aF(k) + ( ::) G~k)]

CL2 = ,::)

and F(k) and G(k) are the real and imaginary parts of the Theodorsen function C(k)plotted in Fig. 5.11 (Theodorsen, 1934). The lift that acts at the rnidchord is propor-tional to a = Zluo, where z is the translation (vertically downward) of the airfoil. Thatis, it represents a force opposing the downward acceleration of the airfoil. This forceis exactly that which is required to impart an acceleration z to a mass of air containedin a cylinder, the diameter of which equals the chord c. This is known as the "appar-ent additional mass." It is as though the mass of the airfoil were increased by thisamount. Except in cases of very low relative density JL = 2m/pSi:, this added mass issmall compared to that of the airplane itself, and hence the force CL2 is relativelyunimportant. Physically, the origin of this force is in the reaction of the air which isassociated with its downward acceleration. The other component, CLI' which acts atthe t chord point, is associated with the circulation around the airfoil, and is a conse-quence of the imposition of the Kutta-Joukowski condition at the trailing edge. It isseen that it contains one term proportional to a and another proportional to a. FromFig. 5.10, the pitching-moment coefficient about the CG is obtained as

where (5.5,6)

(5.5,7)

6Rodden and Giesing (1970) have extended and generalized this method. In particular they give re-sults for finite wings.

Page 156: Stability Control EtkinReid

146 Chapter 5. The Stability Derivatives

1.0r\\

<,~ ...••. I'-....r--. --

0.9

0.8F(i)

0.7

0.6

0.5o ~ ~ M ~ ~ ~ ~ M M W

Reduced frequency, k

0.20

0.04

I-

J~•...•••...

to...

/...•.• '-...

""""- -I --r--I

0.16

0.12-G(k)

0.08

~ ~ ~ M ~ ~ ~ ~ M M WReduced frequency, k

Figure 5.11 The Theodorsen function.too

From [(5.5,6) and (5.5,7)], the following derivatives are found for frequency k.

CLa = 21TF(k)

G(k)CL;" = 17'+ 21T-

k-

Cma = 21TF(k)(h - !)G(k)

Cm;" = 7T(h -i) + 21T-k- (h - t)

••

(5.5,8)

The awkward situation is evident, from (5.5.8), that the derivatives are frequency-de-pendent. That is, in free oscillations one does not know the value of the derivative un-til the solution to the motion (i.e. the frequency) is known. In cases of forced oscilla-tions at a given frequency, this difficulty is not present.

When dealing with the rigid-body motions of flight vehicles, the characteristicnondimensional frequencies k are usually small, k ~ 1. Hence it is reasonable to usethe F(k) and G(k) corresponding to k ~ O. For the two-dimensional incompressiblecase described above, lim F(k) = 1, so that CLa = 217'and Cma = 27T(h- t), the theo-retical steady-flow val~es~This conclusion, that CLa and Cma are the quasistatic values,also holds for finite wings at all Mach numbers. The results for CL;" and Cma are notso clear, however, since lim G(k)/k given above is infinite. This singularity is markedfor the example of two-dimensional flow given above, but is not evident for finitewings at moderate aspect ratio. Miles (1950) indicates that the k log k term responsi-

r\

••

Page 157: Stability Control EtkinReid

5.5 The «Derivatives (CLiit, Cm) 147

ble for the singularity is not significant for aspect ratios less than to, and the numeri-cal calculations of Rodden and Giesing (1970) show no difficulty at values of k aslow as 0.001. Filotas' (1971) solutions for finite wings bear out Miles' contention.Thus for finite wings definite values of CLiit and Cm;. can be associated with small butnonvanishing values of k. The limiting values described above can be obtained from afirst-order-in-frequency analysis of an oscillating wing. To summarize, the a deriva-tives of a wing alone may be computed from the indicial response of lift and pitchingmoment, or from first-order-in-frequency analysis of harmonically plunging wings.

CONTRIBUTIONS OF A TAIL

There is an approximate method for evaluating the contributions of a tail surface,which is satisfactory in many cases. This is based on the concept of the lag of thedownwash. It neglects entirely the nonstationary character of the lift response of thetail to changes in tail angle of attack, and attributes the result entirely to the fact thatthe downwash at the tail does not respond instantaneously to changes in wing angleof attack. The downwash is assumed to be dependent primarily on the strength of thewing's trailing vortices in the neighborhood of the tail. Since the vorticity is con-vected with the stream, then a change in the circulation at the wing will not be felt asa change in downwash at the tail until a time At = l/uo has elapsed, where It is thetail length. It is therefore assumed that the instantaneous downwash at the tail, €(t),corresponds to the wing a at time (t - At). The corrections to the quasi static down-wash and tail angle of attack are therefore

de de ~Ae= --il.At= --il.-da da Uo

= -Aat (5.5,9)

CZa ofa TailThe correction to the tail lift coefficient for the downwash lag is

It deAC£, = a,ti.at= atil.--

Uo daThe correction to the airplane lift is therefore

. it de S,AC = aa----

L t Uo da S

(5.5, to)

Therefore

•sc, dCL de It s,--= ---= -a ---dil. dil. t da Uo S

and

(5.5,11)

..

Page 158: Stability Control EtkinReid

148 Chapter 5. The Stability Derivatives

Cmaofa TailThe correction to the pitching moment is obtained from !i.CLr as

de Itsc; = -VH!i.Cz., = -arf;t:;-- VHoa Uo

Therefore

and

(5.5,12)

5.6 The fJ Derivatives (CyfJ, C1fJ, CnfJ)

These derivatives all are obtainable from wind-tunnel tests on yawed models (Camp-bell and McKinney, 1952). Generally speaking, estimation methods do not give com-pletely reliable results, and testing is a necessity.

THE DERIVATIVE CYfJ

This is the side-force derivative, giving the force that acts in the y direction (right)when the airplane has a positive {3or v (i.e., a sideslip to the right, see Fig. 3.11). CYflis usually negative, and frequently small enough to be neglected entirely. The maincontributions are those of the body and the vertical tail, although the wing, and wing-body interference, may modify it significantly. Of these, only the tail effect is readilyestimated. It may be expressed in terms of the vertical-tail lift-curve slope and thesidewash factor (see Sec. 3.9). (In this and the following sections the fin velocity ra-tio VFIV is assumed to be unity.)

or

(dU) SF

(CY/l)tail = -aF 1 - d{3 S (5.6,1)

The most troublesome component of this equation is the sidewash derivative duld{3,which is difficult to estimate because of its dependence on the wing and fuselagegeometry (see Sec. 3.9).

THE DERIVATIVE C'fJ

C1fl is the dihedral effect, which was discussed at some length in Sec. 3.12.

•\

Page 159: Stability Control EtkinReid

THE DERIVATIVE Cnp

Cnp is the weathercock stability derivative, dealt with in Sec. 3.9.

5.7 The p Derivatives «i; Cl", Cn,,)When an airplane rolls with angular velocity p about its x axis (the flight direction),its motion is instantaneously like that of a screw. This motion affects the airflow (lo-cal angle of attack) at all stations of the wing and tail surfaces. This is illustrated inFig. 5.12 for two points: a wing tip and the fin tip. It should be noted that the nondi-mensional rate of roll, ft = pb/2uo is, for small p, the angle (in radians) of the helixtraced by the wing tip. These angle of attack changes bring about alterations in theaerodynamic load distribution over the surfaces, and thereby introduce perturbationsin the forces and moments. The change in the wing load distribution also causes amodification to the trailing vortex sheet. The vorticity distribution in it is no longersymmetrical about the x axis, and a sidewash (positive, i.e., to the right) is induced ata vertical tail conventionally placed. This further modifies the angle-of-attack distri-bution on the vertical-tail surface. This sidewash due to rolling is characterized by thederivative auriJft. It has been studied theoretically and experimentally by Michael(1952), who has shown its importance in relation to correct estimation of the tail con-tributions to the rolling derivatives. Finally, the helical motion of the wing produces atrailing vortex sheet that is not flat, but helical. For the small rates of roll admissiblein a linear theory, this effect may be neglected with respect to both wing and tailforces.

THE DERIVATIVE Cy"

The side force due to rolling is often negligible. When it is not, the contributions thatneed to be considered are those from the wing? and from the vertical tail. The verti-cal-tail effect may be estimated in the light of its angle-of-attack change (see Fig.5.12) as follows. Let the mean change in aF (see Fig. 3.12) due to the rolling velocitybe

PZF audaF= --+p-

Uo apwhere ZF is an appropriate mean height of the fin. Introducing the nondimensionalrate of roll, we may rewrite this as

(ZF au)

daF= -ft 2b- aft

The incremental side-force coefficient on the fin is obtained from daF'

(5.7,1)

A (ZF au)dC = a da = -a..n 2 - - -YF F F /'1' b aft (5.7,2)

7Por the effect of the wing at low speeds, see Campbell and McKinney (1952).

Page 160: Stability Control EtkinReid

150 Chapter 5. The Stability Derivatives

Figure 5.12 Angle of attack changes due to p.

where aF is the lift-curve slope of the vertical tail. The incremental side force on theairplane is then given by

SF A SF (ZF au)dC =-dC =-ap- 2---Y S YF F: S b ap •••

thus

(5.7,3)

THE DERIVATIVE C,p

C, is known as the damping-in-roll derivative. It expresses the resistance of the air-p

plane to rolling. Except in unusual circumstances, only the wing contributes signifi-cantly to this derivative. As can be seen from Fig. 5.12, the angle of attack due to pvaries linearly across the span, from the value pb/2uo at the right wing tip to -pb/2uoat the left tip. This antisymmetric a distribution produces an anti-symmetric incre-ment in the lift distribution as shown in Fig. 5.13. In the linear range this is superim-posed on the symmetric lift distribution associated with the wing angle of attack inundisturbed flight. The large rolling moment L produced by this lift distribution isproportional to the tip angle of attackp (see Fig. 5.12), and Clp is a negative constant,so long as the local angle of attack remains below the local stalling angle.

If the wing angle of attack at the center line, aw(O), is large, then the incrementalvalue due to p may take some sections of the wing beyond the stalling angle, as

••

Page 161: Stability Control EtkinReid

••

II

Figure 5.13 Spanwise lift distribution due to rolling.

shown in Fig. 5.14. [Actually, for finite span wings, there is an additional induced an-gle of attack distribution aly) due to the vortex wake that modifies the net sectionalvalue still further. We neglect this correction here in the interest of making the mainpoint.] When this happens ICI))I is reduced in magnitude from the linear value and ifaw(O) is large enough, will even change sign. When this happens, the wing will au-torotate, the main characteristic of spinning flight.

THE DERIVATIVE c,l'

The yawing moment produced by the rolling motion is one of the so-called cross de-rivatives. It is the existence of these cross derivatives that causes the rolling and yaw-ing motions to be so closely coupled. The wing and tail both contribute to C; .

p

The wing contribution is in two parts. The first comes from the change in profiledrag associated with the change in wing angle of attack. The wing a is increased onthe right-hand side and decreased on the left-hand side. These changes will normallybe accompanied by an increase in profile drag on the right side, and a decrease on the

o awC-f) awCO) awCf)Net section angle of attack

Figure 5.14 Reduction of Cll' due to wing stall.

Page 162: Stability Control EtkinReid

152 Chapter 5. The Stability Derivatives

,;.

y

Figure 5.15 Inclination of CL vector due to rolling.

left side, combining to produce a positive (nose-right) yawing moment. The secondwing effect is associated with the fore-and-aft inclination of the lift vector caused bythe rolling in subsonic flight and in supersonic flight when the leading edge is sub-sonic. It depends on the leading-edge suction. The physical situation is illustrated inFig. 5.15. The directions of motion of two typical wing elements are shown inclinedby the angles ± ()= Py/uo from the direction of the vector Uo' Since the local lift isperpendicular to the local relative wind, then the lift vector on the right half of thewing is inclined forward, and that on the left half backward. The result is a negativeyawing couple, proportional to the product CzP. If the wing leading edges are super-sonic, then the leading-edge suction is not present, and the local force remains nor-mal to the surface. The increased angle of attack on the right side causes an increasein this normal force there, while the opposite happens on the left side. The result is apositive yawing couple proportional to p.

The tail contribution to Cnp is easily found from the tail side force given previ-ously (5.7,2). The incremental C; is given by

SF IF(dCn)tail = - dCYF Sb

where IF is the distance shown in Fig. 3.12. Therefore

SF IF (ZF au)(dCn)taiI = aFfJ Sb 2b- ap

•••

'.

and (5.7,4)

( ZF aU)(Cn)tail = aFVV 2b- ap

where Vv is the vertical-tail volume ratio.

..

Page 163: Stability Control EtkinReid

5.8 The r Derivatives (CYr' c., Cn,) 153••

5.8 The r Derivatives (Cyr, C1r, Cn)When an airplane has a rate of yaw r superimposed on the foward motion uo, its ve-locity field is altered significantly. This is illustrated for the wing and vertical tail inFig. 5.16. The situation on the wing is clearly very complicated when it has muchsweepback. The main feature however, is that the velocity of the t chord line normalto itself is increased by the yawing on the left-hand side, and decreased on the rightside. The aerodynamic forces at each section (lift, drag, moment) are therefore in-creased on the left-hand side, and decreased on the right-hand side. As in the case ofthe rolling wing, the unsymmetrical lift distribution leads to an unsymmetrical trail-ing vortex sheet, and hence a sidewash at the tail. The incremental tail angle of attackis then

rlF aO"AaF=-+r-

Uo dr

or

(5.8,1)

"0

"0

1\f \I1 "0IIIIII1

B

I-----r~-----y

Figure 5.16 Velocity field due to yawing. AB = velocity vector due to rate of yaw r.

Page 164: Stability Control EtkinReid

154 Chapter 5. The Stability Derivatives

THE DERIVATIVE CYr

The only contribution to CYr that is normally important is that of the tail. From the an-gle of attack change we find the incremental C; to be

A SF (iF aCT)(aCY)tail = aFrS 2b + ar

thus

(5.8,2)

THE DERIVATIVE CZr

This is another important cross derivative; the rolling moment due to yawing. The in-crease in lift on the left wing, and the decrease on the right wing combine to producea positive rolling moment proportional to the original lift coefficient CL' Hence thisderivative is largest at low speed. Aspect ratio, taper ratio, and sweepback are all im-portant parameters.

When the vertical tail is large, its contribution may be significant. A formula forit can be derived in the same way as for the previous tail contributions, with the result

SF ZF (iF aU)(Cz)tail=aFSb 2b+ ar (5.8,3)

THE DERIVATIVE c;Cnr is the damping-in-yaw derivative, and is always negative. The body adds a negli-gible amount to Cnr except when it is very large. The important contributions arethose of the wing and tail. The increases in both the profile and induced drag on theleft wing and the decreases on the right wing give a negative yawing moment andhence a resistance to the motion. The magnitude of the effect depends on the aspectratio, taper ratio, and sweepback. For extremely large sweepback, of the order of 600

,

the yawing moment associated with the induced drag may be positive; that is, pro-duce a reduction in the damping.

The side force on the tail also provides a negative yawing moment. The calcula-tion is similar to that for the preceding tail contributions, with the result

(5.8,4)

5.9 Summary of the FormulasThe formulas that are frequently wanted for reference are collected in Tables 5.1 and5.2. Where an entry in the table shows only a tail contribution, it is not implied thatthe wing and body effects are not important, but only that no convenient formula isavailable.

Page 165: Stability Control EtkinReid

••

TableS.lSummary-Longitudinal Derivatives

Cx c, Cm

at (dCT dCD) 2dCD (dCD) dCL 20CL dCL ac, 2 ec; ec;Me aM - dM - P"o 0Pd + CT. 1 - aCT -Mo aM - PUo apd - CT. aCT Mo dM + PUo apd + CT. aCT

a C/o - CD" -(CL" + COo) -a(hn - h)

dE I, dEIx Neg. *-2aV - *-2aV --'H da t He da

q Neg. *-2a,VHI,

*-2aV -t He

Neg. means usually negligible.

*means contribution of the tail only, formula for wing-body not available.

(i'mau)o .tCT. = ~PUoS - 2CTO; CTo= CDo + Cwosm 80

Page 166: Stability Control EtkinReid

156 Chapter 5. The Stability Derivatives

Table 5.2Summary-Lateral Derivatives

Cy C/ Cn

f3 *-a SF (1 _ aU) N.A. *aFvv( 1 - ~;)F S af3p *-a SF (2ZF _ au) N.A. (ZF aU)*a V 2---F S b ap F v b ap

f SF (IF au) SF ZF (IF au) (IF au)*a - 2-+- *a -- 2-+- *-a V 2-+-F S b af FSb b of F v b of

"

*means contribution of the tail only, formula for wing-body not available; VFIV = l.N.A. means no formula available.

5.10 Aeroelastic DerivativesIn Sec. 4.12 there were introduced aerodynamic derivatives associated with the defor-mations of the airplane. These are of two kinds: those that appear in the rigid-bodyequations and those that appear in the added equations of the elastic degrees of free-dom. These are illustrated in this section by consideration of the hypothetical vibra-tion mode shown in Fig. 5.17. In this mode it is assumed that the fuselage and tail arerigid, and have a motion of vertical translation only. The flexibility is all in the wing,and it bends without twisting. The functions describing the mode (4.11,1) are there-fore:

x' = 0y' = 0

z' = h(Y)ZT

For the generalized coordinate, we have used the wing-tip deflection Zr- h(y) is then anormalized function describing the wing bending mode.

Since the elastic degrees of freedom are only important in relation to stabilityand control when their frequencies are relatively low, approaching those of the rigid-body modes, then it is reasonable to use the same approximation for the aerodynamicforces as is used in calculating stability derivatives. That is, if quasisteady flow the-ory is adequate for the aerodynamic forces associated with the rigid-body motions,then we may use the same theory for the elastic motions.

(5.10,1)

-----~C. G. of daformed airplane ....t.

1Figure 5.17 Symmetrical wing bending.

Page 167: Stability Control EtkinReid

5.10 Aeroelastic Derivatives 157

In the example chosen, we assume that the only significant forces are those onthe wing and tail, and that these are to be computed from quasisteady flow theory. Inthe light of these assumptions, some of the representative derivatives of both typesare discussed below. As a preliminary, the forces induced on the wing and tail by theelastic motion are treated first.

FORCES ON THE WINGThe vertical velocity of the wing section distant y from the center line is

Z = h(Y)zT

and the corresponding change in wing angle of attack is

Lla(y) = h(Y)zTluo

(5.10,2)

(5.10,3)

This angle of attack distribution can be used with any applicable steady-flow wingtheory to calculate the incremental local section lift. (It will of course be proportionalto iluo.) Let it be denoted in coefficient form by C;(y)zTluo, and the correspondingincrement in wing total lift coefficient by C'z-.,ZT1uO'C;(y) and C'z-., are thus the valuescorresponding to unit value of the nondimensional quantity ZT1uO'

FORCE ON THE TAIL

The tail experiences a downward velocity h(O)zT' and also, because of the alteredwing lift distribution, a downwash change (d€ldzT)ZT' Hence the net change in tail an-gle of attack is

dELlat = h(O)zT1uo - -d' ZT

ZT

[dE] ZT= h(O) - -

d(ZT!UO) Uo

This produces an increment in the tail lift coefficient of amount

[dE] ZT

LlCL, = at h(O) - d(zTluo

) ~~ (5.10,4)

THE DERIVATIVE ZiT

This derivative describes the contribution of wing bending velocity to the Z force act-ing on the airplane. A suitable nondimensional form is dC/d(ZT!UO)' Since Cz = -CLl

we have that

, ZT [ dE J ZTLlC = -C - - a h(O) - -z Lw Uo t d(ZT!UO) Uo

and hence

(5.10,5)

Page 168: Stability Control EtkinReid

158 Chapter 5. The Stability Derivatives

THE DERIVATIVE Anw

This derivative [see (4.12,12)] represents the contribution to the generalized force inthe bending degree of freedom, associated with a change in the w velocity of the air-plane. A suitable nondimensional form is obtained by defining

?}

C!!F= \pu~

and using ax in place of w (w = uoax)' Then the appropriate nondimensional deriva-tive is C!!Fa'

Let the wing lift distribution due to a perturbation a in the angle of attack (con-stant across the span) be given by C1a(y)a. Then in a virtual displacement in the wingbending mode 5zn the work done by this wing loading is

rbl2 15W = - J aC1a(y)h(y) 5zT-2 pu~c(y) dy

-b12

where c(y) is the local wing chord. The corresponding contribution to?} is

5W 1 rb12

-a = -a -2 pu~ J C1a(y)h(y)c(y) dyZT -b12

and to C!!Fais

1 a2w 1 rb12

1 2", -a ::l = - -S J C1a(y)h(y)c(y) dy2PUQiJ z']Va -b12

The tail also contributes to this derivative, for the tail lift associated with a is

(5.10,6)

( aE) 1 2a.a 1 - aa "2 pUoSt

and the work done by this force during the virtual displacement is

Therefore the contribution to C!!Fis

s, ( aE)-atsh(O) 1 - aaThe total value of C!!Fais then the sum of 5.10,6 and 5.10,7.

(5.10,7)

THE DERIVATIVE bu (see 4.12,12)

This derivative identifies the contribution of ZT to the generalized aerodynamic forcein the distortion degree of freedom. We have defined the associated wing load distri-

Page 169: Stability Control EtkinReid

5.11 Exercises 159

bution above by the local lift coefficient q(Y)ZT1uO' As in the case of the derivativeAnw above, the work done by this loading is calculated, with the result that the wingcontributes

«; I d2W I fbl2-- = -- = - - C'(y)h(y)c(y) dyd(zTluo) ipu~ dZTd(ZT1uO) S -ea 1

Likewise, the contribution of the tail is calculated here as for Anw' and is found to be

(5.10,8)

~ [ dE]-at Sh(O) h(O) - d(ZT1uO)

The total value of dC~/a(ZTluo) is then the sum of 5.10,8 and 5.10,9.

(5.10,9)

5.11 Exercises

5.1 The derivative c; contains the term Mo (~~ t Estimate the magnitude of this term

for an airplane with wing loading 70 psf (3,352 Pa) flying at 20,000 ft (6,096 m) alti-tude, for Mach numbers between 0.2 and 0.8. The following data pertain to the wing:

Sweep (i chord) A = 30°

S = 5,500 ft2 (511.0 nr')

Plot the result vs. Mo. Calculate the contribution this term makes to Z; and plot thisas well. (Compare with Zu for the B747 from Table 6.2, and comment).

5.2 A wind-tunnel model is mounted with one degree of freedom-pivoted so that it canonly rotate about the y-axis of the body frame, which is perpendicular to the relativewind. It is elastically restrained with a pitching moment M = - kO. Show how thesum (Cmq + Cm,,) can be estimated from experiments in which the model is free to os-cillate in pitch with wind on and off. Assume Mq can be neglected with the wind offand Zq and Zw can be neglected.

5.3 Consider the wind/fin system of Fig. 5.16, with the following properties:

Wing: A = 5; A = 0.5; A1I4 = 30°; r variable

Fin: aF = 3.5 rad""; IFlb = 0.5; zFlb = 0.1; Vvvariable; au/af3negligible.

Estimate values of the stability derivatives (for hnw = h and LID = 12)

at CLo = 1.0. Plot the spiral stability boundary for horizontal flight:

E = C1pCn, - C1,Cnp= 0[see (6.8,6) with (}o = 0] in the plane of Vv vs. r. (Make any reasonable assumptionsyou need to supplement the given data).

5.4 A jet airplane has a thrust line that passes above the CO by a distance equal to 10%of the M.A.C. With the assumption aTldU = 0, estimate the increment thus caused inc.;

Page 170: Stability Control EtkinReid

160 Chapter 5. The Stability Derivatives

5.5 Find C; due to the tilting of the lift vector for a wing with an elliptic lift distributionp f 2

(i.e., a wing with lift per unit span l(y) which obeys a2 + ~ = 1)- Assume that the

tilt angle is small. Express Cnp in terms of Ct» the lift coefficient of the wing when it isnot rolling.

5.6 Assume that Figs. 5.7 and 5.8 are experimental measurements. Select an analyticfunction C4tep(t) that can represent Fig. 5.7 (M = 0 case). Find the correspondingtransfer function relating CL to a. Use this transfer function to generate a function oftime corresponding to Fig. 5.8b and demonstrate that it has the desired form.

5.12 Additional Symbols Introduced in Chapter 5

e efficiency factor for wing-induced drag

k reduced frequency, we/2uoPd dynamic pressure, ipy2co circular frequency

See also Sees. 2.11, 3.15, and 4.14.

"

Page 171: Stability Control EtkinReid

CHAPTER 6

Stability of UncontrolledMotion

The preceding chapters have been to some extent simply the preparation for what fol-lows in this and the succeeding two chapters, that is, a treatment of the uncontrolledand controlled motions of an airplane. The system model was developed in Chap. 4,and the aerodynamic ingredients were described in Chaps. 2, 3, and 5. In this chapterwe tackle first the simplest of these cases, the uncontrolled motion, that is, the motionwhen all the controls are locked in position. An airplane in steady flight may be sub-jected to a momentary disturbance by a nonuniform or nonstationary atmosphere, orby movements of passengers, release of stores, and so forth. In this circumstancesome of the questions to be answered are, "What is the character of the motion fol-lowing the cessation of the disturbance? Does it subside or increase? If it subsideswhat is the final flight path?" The stability of small disturbances from steady flight isan extremely important property of aircraft-first, because steady flight conditionsmake up most of the flight time of airplanes, and second, because the disturbances inthis condition must be small for a satisfactory vehicle. If they were not it would beunacceptable for either commercial or military use. The required dynamic behavior isensured by design-by making the small-disturbance properties of concern (the nat-ural modes) such that either human or automatic control can keep the disturbances toan acceptably small level. Finally the small-disturbance model is actually valid fordisturbance magnitudes that seem quite violent to human occupants.

6.1 Form of Solution of Small-Disturbance Equations

The small-disturbance equations are (4.9,18 and 4.9,19). They are both of the form

t = Ax + afc (6.1,1)

where x is the (N Xl) state vector, A is the (N X N) system matrix, a constant, andMc is the (N Xl) vector of incremental control forces and moments. In this applica-tion, the control force vector is zero, so the equation to be studied is

t=Ax (6.1,2)

Solutions of this first-order differential equation are well known. They are of theform

(6.1,3)

Xo is an eigenvector and A is an eigenvalue of the system. Xo is also seen to be thevalue of the state vector at t = O.Substitution of (6.1,3) into (6.1,2) gives

161

Page 172: Stability Control EtkinReid

162 Chapter6. Stability of UncontrolledMotion

orAxo = AXo

(A - AI)Xo = 0(6.1,4)(6.1,5)

where I is the identity matrix. Since the scalar expansion of (6.1,5) is a system of Nhomogeneous equations (zeros on the right-hand side) then there is a nonzero solu-tion for Xo only when the system determinant vanishes, that is, when

det (A - AI) = 0 (6.1,6)

The determinant in (6.1,6) is the characteristic determinant of the system. When ex-panded, the result is a polynomial in A of degree N, the characteristic polynomial,and the Nth degree algebraic equation (6.1,6) is the characteristic equation of thesystem. Since the equation is of the Nth degree it has in general N roots Ai' some realand some occurring in conjugate complex pairs. Corresponding to each real eigen-value A is a real eigenvector Xo, and to each complex pair Ai and A; there correspondsa conjugate complex pair of eigenvectors Xo and ~. Since anyone of the A'S can pro-vide a solution to (6.1,2) and since the equation is linear, the most general solution isa sum of all the corresponding x(t) of (6.1,3), that is,

x(t) =L Xo,eA,ti

(6.1,7)

Each of the solutions described by (6.1,3) is called a natural mode, and the generalsolution (6.1,7) is a sum of all the modes. A typical variable, say w, would, accordingto (6.1,7) have the form

(6.1,8)

where the a, would be fixed by the initial conditions. The pair of terms correspondingto a conjugate pair of eigenvalues

A = n ± it» (6.1,9)(6.1,10)is

Upon expanding the exponentials, (6.1,10) becomes

ent(AI cos wt + A2 sin wt) (6.1,11)

where Al = (al + a2) and A2 = ita, - a2) are always real. That is, (6.1,11) describesan oscillatory mode, of period T = 21rlw, that either grows or decays, depending onthe sign of n. The four kinds of mode that can occur, according to whether Ais real orcomplex, and according to the sign of n are illustrated in Fig. 6.1. The disturbancesshown in (a) and (c) increase with time, and hence these are unstable modes. It isconventional to refer to (a) as a static instability or divergence, since there is no ten-dency for the disturbance to diminish. By contrast, (c) is called dynamic instability ora divergent oscillation, since the disturbance quantity alternately increases and di-minishes, the amplitude growing with time. (b) illustrates a subsidence or conver-gence, and (d) a damped or convergent oscillation. Since in both (b) and (d) the dis-turbance quantity ultimately vanishes, they represent stable modes.

It is seen that a ''yes'' or "no" evaluation of the stability is obtained simply fromthe signs of the real parts of the AS. If there are no positive real parts, there is no in-stability. This information is not sufficient, however, to evaluate the handling quali-

Page 173: Stability Control EtkinReid

6.1 Form of Solution of Small-Disturbance Equations 163

(s)

"oL..----L.----- __tdoubl•

Ie)

..••...••....... ..

II

3c:'iii

••~+3§.;:\:o L..- l....- t

tho"

Ib)

/

Id)

Figure 6.1 Types of solution. (a) Areal, positive. (b) Areal, negative. (c) Acomplex, n > O. (d) Acomplex, n < O.

ties of an airplane (see Chap. 1). These are dependent on the quantitative as well ason the qualitative characteristics of the modes. The numerical parameters of primaryinterest are

21T1. Period, T = -

to2. Time to double or time to half.3. Cycles to double (Ndouble) or cycles to half (Nhalf).

The first two of these are illustrated in Fig. 6.1. When the roots are real, there is ofcourse no period, and the only parameter is the time to double or half. These are thetimes that must elapse during which any disturbance quantity will double or halve it-self, respectively. When the modes are oscillatory, it is the envelope ordinate thatdoubles or halves. Since the envelope may be regarded as an amplitude modulation,

Page 174: Stability Control EtkinReid

164 Chapter 6. Stability of Uncontrolled Motion

the~ we may think of the doubling or halving as applied to the variable amplitude. By~otmg that loge 2 = -loge i = 0.693, the reader will easily verify the following rela-nons:

Time to double or half:

.693 .693tdoubleor thalf = ¥ = WW

n

Cycles to double or half:

W ~Ndoubleor Nhalf = .1l0 ~ = .110 III

Logarithmic decrement (log of ratio of successive peaks):

ent l8 = loge n(t+T) = -nT = 271' V Z

e 1 -l

(a)

(b) (6.1,12)

= - .693/Ndouble or .693/Nhalf (c)

In the preceding equations,

to; = (wz + nZ) 1/2, the "undamped" circular frequencyl = - n/ Wno the damping ratio

COMPUTATION OF EIGENVALUES AND EIGENVECTORS

As noted above, the eigenvalues and eigenvectors are properties of the matrix A. Anumber of software packages such as MATLAB are now available for calculatingthem. For all the numerical examples in this book we have used the Student Versionof Program cc. 1Appendix A.5 shows how we used it to get the results.

ROUTH'S CRITERIA FOR STABILITY

The stability of the airplane is governed, as we have seen, by the real parts of theeigenvalues, the roots of the characteristic equation. Now it is not necessary actuallyto solve the characteristic equation (6.1,6) for these roots in order to discover whetherthere are any unstable ones. E. J. Routh (1905) has derived a criterion that can be ap-plied to the coefficients of the equation to get the desired result. The criterion is that acertain set of test functions shall all be positive (Etkin, 1972). We present below theresult for the important case of the quartic equation, which will turn up later in thischapter.

Let the quartic equation be

AA4 + BA3 + CAz + DA + E = 0 (A > 0) (6.1,13)

Then the test functions are Fo = A, F1 = B, Fz = BC - AD, F3 = FzD - BZE, F4 =

rI

1Available from Systems Technology Inc. Hawthorne, CA.

Page 175: Stability Control EtkinReid

6.2 Longitudinal Modes of a Jet Transport 165

F3BE. The necessary and sufficient conditions for these test functions to be positiveare

A,B,D,E>O

and

R = D(BC - AD) - B2E> 0 (6.1,14)

It follows that C also must be positive. The quantity on the left-hand side of (6.1,14)is commonly known as Routh's discriminant.

Duncan (1952) has shown that the vanishing of E and of R represent significantcritical cases. If the airplane is stable, and some design parameter is then varied insuch a way as to lead to instability, then the following conditions hold:

1. If only E changes from + to -, then one real root changes from negative topositive; that is, one divergence appears in the solution (see Fig. 6.1a).

2. If only R changes from + to -, then the real part of one complex pair of rootschanges from negative to positive; that is, one divergent oscillation appears inthe solution (see Fig. 6.1c).

Thus the conditions E = 0 and R = 0 define boundaries between stability and in-stability. The former is the boundary between stability and static instability, and thelatter is the boundary between stability and a divergent oscillation. These stabilityboundaries are very useful for certain analytical purposes. We shall in particularmake use of the E = 0 boundary.

6.2 Longitudinal Modes of a Jet TransportThe foregoing theory is now illustrated by applying it to the Boeing 747 transport.The needed geometrical and aerodynamic data for this airplane are given in AppendixE. The flight condition for this example is cruising in horizontal flight at approxi-mately 40,000 ft at Mach number 0.8. Relevant data are as follows:

I

IW = 636,636 lb (2.83176 X 106 N)C = 27.31 ft (8.324 m)

I; = 0.183 X 108 slug ft2 (0.247 X 108 kg rrr')

CLo = 0.654

S = 5500 ft2 (511.0 m2)

b = 195.7 ft (59.64 m)

Iy = 0.331 X 108 slug fe(0.449 X 108 kg m2)

Izx = - .156 X 107 slug ft2(- .212 X 107 kg nr')

p = 0.0005909 slug/ft"(0.3045 kg/m")

CDo = 0.0430

uo = 774 fps (235.9 m/s) 00 = 0

The preceding four inertias are for stability axes at the stated flight condition. In thenumerical examples of this and the following two chapters, the system matrices andthe solutions are all given in English units. The nondimensional stability derivatives

Page 176: Stability Control EtkinReid

166 Chapter 6. Stability of Uncontrolled Motion

Table 6.1Nondimensional Derivatives-B747 Airplane

Cx Cz CmU -0.1080 -0.1060 0.1043a 0.2193 -4.920 -1.023q 0 -5.921 -23.92~ 0 5.896 -6.314a

are given in Table 6.1, and the dimensional derivatives in Table 6.2. With the abovedata we calculate the system matrix A for this case. (Recall that the state vector is[du w q MJ]1). [

-0.006868 0.01395 0 -32'~02]-0.09055 -0.3151 773.98

A = 0.0001187 -0.001026 -0.4285 (6.2,1)001

The characteristic equation (6.1,6) is next calculated to be:

A4 + 0.750468".3 + 0.935494A2 + 0.0094630A + 0.0041959 = 0 (6.2,2)

The two stability criteria are

andE = 0.0041959 > 0R = 0.004191 > 0

so that there are no unstable modes.

EIGENVALUES

The roots of the characteristic equation (6.2,2), the eigenvalues, areMode 1 (Phugoid mode): A1•2 = -0.003289 ± 0.06723iMode 2 (Short-period mode): A3,4= -0.3719 ± 0.8875i

We see that the natural modes are two damped oscillations, one of long period andlightly damped, the other of short period and heavily damped. This result is quite typ-

(6.2,3)

Table 6.2Dimensional Derivatives-B747 Airplane

X(lb) Z(lb) M (ft-lb}

u(ftIs) -1.358 X 102 -1.778 X 103 3.581 X 1&w(ft/s) 2.758 X 102 -6.188 X 103 -3.515 X litq(rad/s) 0 -1.017 X lOS -1.122 X 107

w(ftIs2) 0 1.308 X 1Q2 -3.826 X 103

X(N) Z(N) M(m'N)

u(mls) -1.982 X 1& -2.595 X 104 1.593 X 104

w(m1s) 4.025 X 103 -9.030 X 104 -1.563 X litq(rad/s) 0 -4.524 X lOS -1.521 X 107

W(m1s2) 0 1.909 X 1& -1.702 X lit

Page 177: Stability Control EtkinReid

6.2 Longitudinal Modes of a Jet Transport 167

Table 6.3

Period thalf Nhalf

Mode Name (s) (s) (cycles)

1 Phugoid" 93.4 211 22.52 Short-period 7.08 1.86 0.26

"The phugoid mode was first described by Lanchester (1908), who also named it. The namecomes from the Greek root for flee as iu fugitive. Actually Lanchester wanted the root forfly. Appropriate or not, the word phugoid has become established in aeronautical jargon.

ical. The modes are conventionally named as in Table 6.3, which also gives their peri-ods and damping. The transient behavior of the state variables in these two modes isdisplayed in Fig. 6.2.

EIGENVECTORS

The eigenvectors corresponding to the above modes are given in Table 6.4. They arefor the nondimensionable variables, in polar form, the values given corresponding ton + iw. Eigenvectors are arbitrary to within a complex factor, so it is only the relativevalues of the state variables that are significant. We have therefore factored them to

2,-----------------------,1.5

B 747 Phugoid mode

§ .5'';::;.e a'"t&. -.5

-1

-1.5

-2 '-- __ --,-'--,--__ -:-'- ...J.... ...L-__ ----Ja 600

Time,s

(a)

1.2

.8

" .60.~e .4'"tQl .2c,

a-.2

B 747 Short-period mode

12 18Time,s

(b)

Figure 6.2 Characteristic transients. (a) Phugoid mode. (b) Short-period (pitching) mode.

Page 178: Stability Control EtkinReid

168 Chapter 6. Stability of Uncontrolled Motion

Table 6.4Eigenvectors (polar form)

Phugoid Short-Period

Magnitude Phase Magnitude Phase

au 0.62 92.4° 0.029 57.4°a=w 0.036 82.8° 1.08 19.2°

q 0.0012 92.8° 0.017 112.7°M 1.0 0° 1.0 0°

make d() equal to unity, as displayed in the Argand diagram of Fig. 6.3. As we havechosen the positive value of w, the diagrams can be imagined to be rotating counter-clockwise and shrinking, with their projections on the real axis being the real valuesof the variables.

The phugoid is seen to be a motion in which the pitch rate q and the angle of at-tack change a are very small, but da and d() are present with significant magnitude.The speed leads d()by about 90° in phase.

The short-period mode, by contrast, is one in which there is negligible speedvariation, while the angle of attack oscillates with an amplitude and phase not muchdifferent from that of d(). This mode behaves like one with only two degrees of free-dom, d() and a.

FLIGHT PATHS IN THE CHARACTERISTIC MODES

Additional insight into the modes is gained by studying the flight path. With the at-mosphere at rest, the differential equations for the position of the CG in FE are givenby (4.9,10), with ()o = 0, that is,

,

axE = dUdtE = -Uo() + w

In a characteristic oscillatory mode with eigenvalues A, A*, the variations of dU, (),and w are [cf. (6.1,8)]

(6.2,4)

du = uljeAt + uijeA*t

W = UzjeAt + u~jeA*t

()= u4jeAt + u:jeA*t

where the constants uij are the components of the eigenvector corresponding to A. Forthe previous numerical example, they are the complex numbers given in polar form inTable 6.4. After substituting (6.2,5) in (6.2,4) and integrating from t = 0 to t we get

(6.2,5)

*U1· U u *XE = Uot + -{eAt + A: eA t + const

[Ul' ]= uot + Ze" Re -{ eiwt + const (6.2,6)

Page 179: Stability Control EtkinReid

6.2 Longitudinal Modes of a Jet Transport 169

1ma = ~ = 0.036(not visible)~ = 0.0012 \(not visible)

Jill = 1.0---------{.==l=========:> ~ Reo

(a)

1m

~ = 0.017 Jifl = 0.029(not visible) (not visible)

Jill = 1.0--------~~=l======~> ~Re

a = ~ = 1.08

(b)

Figure 6.3 (a) Vector diagram ofphugoid mode. (b) Vector diagram of short-period mode.

where Re denotes the real part of the complex number in the square brackets. For thenumerical data of the above example (6.2,6) has been used to calculate the flightpaths in the two modes, plotted in Fig. 6.4. The nonzero initial conditions are arbi-trary, and the trajectories for both modes asymptote to the steady reference flightpath. Figure 6.4 shows that the phugoid is an undulating flight of very long wave-length. The mode diagram, Fig. 6.3a shows that the speed leads the pitch angle byabout 90°, from which we can infer that u is largest near the bottom of the wave andleast near the top. This variation in speed results in different distances being traversed

Page 180: Stability Control EtkinReid

170 Chapter 6. Stability of Uncontrolled Motion

-1000

-15001------------ -,

1000

500

-400

-200

200

400

-500-

-1,000.--------------------.

500f-

o -------------

1,000'-- .L..-1 ...I...-1---------'2,000 4.000 6,000

o ----

XE ft(c)

Figure 6.4 (a) Phugoid flight path (fixed reference frame). (b) Phugoid flight path (movingreference frame). (c) Short-period flight path.

during the upper and lower halves of the cycle, as shown in Fig. 6.4a. For larger am-plitude oscillations, this lack of symmetry in the oscillation becomes much more pro-nounced (although the linear theory then fails to describe it accurately) until ulti-mately the upper part becomes first a cusp and then a loop (see Miele, 1962, p. 273).The motion (see Sec. 6.3) is approximately one of constant total energy, the rising

Page 181: Stability Control EtkinReid

6.3 Approximate Equations for the Longitudinal Modes 171

and falling corresponding to an exchange between kinetic and potential energy. Fig-ure 6.4b shows the phugoid motion relative to axes moving at the reference speed "0'

This is the relative path that would be seen by an observer flying alongside at speed"0'

Figure 6.4c shows the path for the short-period mode. The disturbance is rapidlydamped. The transient has virtually disappeared within 3000 ft of flight, even thoughthe initial Aa and A (J were very large. The deviation of the path from a straight line issmall, the principal feature of the motion being the rapid rotation in pitch.

6.3 Approximate Equations for the Longitudinal ModesThe numerical solutions for the modes, although they certainly show their properties,do not give much physical insight into their genesis. Now each oscillatory mode isequivalent to some second-order mass-spring-damper system, and each nonoscilla-tory mode is equivalent to some mass-damper system. To understand the modes, andthe influence on them of the main flight and vehicle parameters, it is helpful to knowwhat contributes to the equivalent masses, springs, and dampers. To achieve this re-quires analytical solutions, which are simply not available for the full system ofequations. Hence we are interested in getting approximate analytical solutions, ifthey can reasonably represent the modes. Additionally, approximate models of the in-dividual modes are frequently useful in the design of automatic flight control systems(McRuer et al., 1973). In the following we present some such approximations and themethods of arriving at them.

There are two approaches generally used to arrive at these approximations. Oneis to write out a literal expression for the characteristic equation and, by studying theorder of magnitude of the terms in it, to arrive at approximate linear or quadratic fac-tors. For example, if the characteristic equation (6.1,13) is known to have a "small"real root, an approximation to it may be obtained by neglecting all the higher powersof A, that is,

DA+E=O

Or if there is a "large" complex root,· it may be approximated by keeping only thefirst three terms, that is,

AA2 + BA + C= 0

This method is frequently useful, and is sometimes the only reasonable way to get anapproximation.

The second method, which has the advantage of providing more physical insight,proceeds from a foreknowledge of the modal characteristics to arrive at approximatesystem equations of lower order than the exact ones. For the longitudinal modes weuse the second method (see below), and for the lateral modes (see Sec. 6.8) bothmethods are needed.

It should be noted that no simple analytical approximations can be relied on togive accurate results under all circumstances. Machine solutions of the exact matrixis the only certain way. The value of the approximations is indicated by examples inthe following.

To proceed now to the phugoid and short-period modes, we saw in Fig. 6.3 thatsome state variables are negligibly small in each of the two modes. This fact suggests

Page 182: Stability Control EtkinReid

172 Chapter 6. Stability of Uncontrolled Motion

certain approximations to them based on reduced sets of equations of motion arrivedat by physical reasoning. These approximations, which are quite useful, are devel-oped below.

PHUGOID MODE

Lanchester's original solution (Lanchester, 1908) for the phugoid used the assump-tions that aT = 0, lia == 0 and T - D == 0 (see Fig. 2.1). It follows that there is no netaerodynamic force tangent to the flight path, and hence no work done on the vehicleexcept by gravity. The motion is then one of constant total energy, as suggested previ-ously. This simplification makes it possible to treat the most general case with largedisturbances in speed and flight-path angle (see Miele, 1962, p. 271 et seq.) Here wecontent ourselves with a treatment of only the corresponding small-disturbance case,for comparison with the exact numerical result given earlier. The energy condition is

E = imy2 - mgZE = consty2 = u~ + 2gzE (6.3,1)or

where the origin of FE is so chosen that Y = Uo when ZE = O. With a constant, and inaddition neglecting the effect of q on CL> then CL is constant at the value for steadyhorizontal flight, that is, CL = CLo = Cwo' and L = Cwoipy2S or, in view of (6.3,1),

L = Cwoipu~ + (CwoPgS)ZE = W + kZE (6.3,2)

Thus the lift is seen to vary linearly with the height in such a manner as always todrive the vehicle back to its reference height, the "spring constant" being

(6.3,3)

The equation of motion in the vertical direction is clearly, when T - D = 0, and 'Y =angle of climb (see Fig. 2.1)

W - Leos 'Y = mZE

or for small 'Y,

W-L=mzE

On combining (6.3,2) and (6.3,4) we get

mZE + kZE = 0

which identifies a simple harmonic motion of period

T= 2'7T E = 2'7T ~V k V c;:;:;;gs

(6.3,4)

Since Cwo = mg/~pu'f,S, this becomes

~ ~ UoT= '7Tv2- = 0.138uog

when Uo is in fps, a beautifully simple result, suggesting that the phugoid period de-pends only on the speed of flight, and not at all on the airplane or the altitude! This

(6.3,5)

Page 183: Stability Control EtkinReid

6.3 Approximate EqUlltionsfor the Longitudinal Modes 173

elegant result is not only of historical interest-it actually gives a reasonable approxi-mation to the phugoid period of rigid airplanes at speeds below the onset of signifi-cant compressibility effects. Thus for the B747 example, the Lanchester approxima-tion gives T = 107s, a value not very far from the true 93s. It is possible to get aneven better approximation, one that gives an estimate of the damping as well. Be-cause q is approximately zero in this mode, we can infer that the pitching moment isapproximately zero-that the airplane is in quasistatic pitch equilibrium during themotion. Moreover the pitching moment can reasonably be simplified in these circum-stances by keeping only the first two terms on the right side of (4.9,17e). Because qand w are both relatively small we further neglect Zq and Z,;.as well. On making thesesimplifications to (4.9,18) and setting ~fc = 0 and 00 = 0 we get the reduced systemof equations:

Xu Xw0

[:i]~m mZu Zwm m

Uo

u, Mw 00 0 1

(6.3,6)

These equations are not in the canonical form x = Ax, but we can still get the charac-teristic equation by substituting x = Xoe

At and factoring out the exponential. The re-sult is

(XJm - A)Zu/mu;o

Equation (6.3,7) expands to

x.i« 0(Zw/m - A) Uo

u; 0o I

-goo-A

=0 (6.3,7)

AA2 + BA + C = 0or A2 + 2'wnA + w~ = 0 (6.3,8)

which is a convenient way to write the characteristic equation of a second-order sys-tem. The constants are

(6.3,9)

from which we derive the radian frequency and damping to be

w~ = - ---L (Zu _ MuZw)

mu., u;,= _2-[---L (Mu z; _ Zu)J-1I2[.!.- Mu + ~ (Xu _ Mu Xw)J

2 muo Mw Uo u; m Mw

(6.3,10)

Page 184: Stability Control EtkinReid

174 Chapter 6. Stability ofUncontroUed Motion

When Mu =0, these reduce to:

UoB=-XMm u W

(6.3,11)

gC=-ZMm u W

from which

2 C gZuW =-= ---

n A muo(a)

Xu~,= -2 V-mgZ~

When Zu from Table 4.4 with Czu = 0 is substituted into (6.3,12) we find (see Exer-cise 6.1) that T; = 271'/wn is exactly the Lanchester period (6.3,5). Moreover if wemake the further assumption that the airplane is a jet with constant thrust (dT/dUo =0) we find the damping ratio to be

(b) (6.3,12) ,

1 CDo

'="Vi' CLo

that is, it is proportional to the inverse of the UD ratio of the airplane. In fact the ap-proximation for the period is good over the whole range of Cmu' whereas that for thedamping is poor for large positive Cmu' For the example airplane the above approxi-mation gives' = 0.046, compared with the exact value 0.049.

SHORT-PERIOD MODE

Figure 6.3b shows that the short-period mode is essentially one with two degrees offreedom, the speed being substantially constant while the airplane pitches relativelyrapidly. We can therefore arrive at approximate system equations by neglecting theX-force equation entirely and putting du = O.Examination of the magnitudes of theterms in the numerical example shows that Zw is small compared to m and Zq is smallcompared to mu-: The result after simplifying (4.9,18) with 80 = 0 is a pair of equa-tions for w and q. z,

[ ; ] ~ [ :,[Mw:M~] :, ~,: M~.]r;] (6.3.13)

The characteristic equation of (6.3,13) is found to be

(6.3,14)

Page 185: Stability Control EtkinReid

6.4 General Theory of Static Longitudinal Stability 175

When converted with the aid of Tables 4.1 and 4.4, (6.3,14) becomes

11.2 + BA + C = 0 (a)

where

1 [Cz 1 ]B = - - _a + - (C + C .'t* 2JL t, mq ma/

1 ( CmqCza)C = - t*2fy Cma - 2JL

When the data for the B747 is substituted into (6.3,15) the result obtained is

(b) (6.3,15)

(c)

with roots

11.2+ 0.74111.+ 0.9281 = 0

A = -0.371 ± 0.889i

which are seen to be almost the same as those in (6.2,3) obtained from the completematrix equation. The short-period approximation is actually very good for a widerange of vehicle characteristics and flight conditions.

6.4 General Theory of Static Longitudinal StabilityIn Chap. 2 we used positive pitch stiffness (negative Cm) as an approximate criterionfor static longitudinal stability. Now static instability really means the presence of areal positive root of the characteristic equation, and we saw in Sec. 6.1 that the condi-tion for no such root to occur is that the coefficient E of the stability quartic must bepositive. Thus the boundary between static stability and instability is defined by E =O. We get E by putting A = 0 in the characteristic determinant. Thus

E = detA (6.4,1)

In evaluating (6.4,1) for the matrix of (4.9,18) we put 00 = 0, and as in Sec. 6.3, weneglect the two derivatives Zw and Zq. The result is

sE = - (ZuMw - MuZw)ml;

Since g, m and Iy are all positive, the criterion for static stability is

z.u; - MuZw > O.

When converted to nondimensional form, this becomes

(6.4,2)

(6.4,3)

(6.4,4)

When there are no speed effects, that is, Czu and Cmu are both zero, then the criteriondoes indeed reduce to the simple Cma < O.

We now compare the above criterion for stability with the trim slope (2.4,24). Inmaking this comparison, we must take note of a minor difference in basic assump-tions. In the preceding development, it was specifically assumed that the thrust vectorrotates with the vehicle when a is changed [see (5.1,1)]. In the development leading

Page 186: Stability Control EtkinReid

176 Chapter 6. Stability of Uncontrolled Motion

to (2.4,24) by contrast, there is an implicit assumption that the thrust provides nocomponent of force perpendicular to V [see (2.4,18)]. It is this difference that leads tothe presence of CZa in (6.4,4) instead of CLa in (2.4,24). Had the assumptions been thesame, the expressions would be strictly compatible. In any case, CDo is usually smallcompared to CLa' so that the difference is not important, see Table 5.1. We see thatthe justification for the statement made in Sec. 2.4, that the slope of the elevator trimcurve (d8etrimld~l5p is a criterion of static stability, is provided by (6.4,4). [Note thatCwo = CLo in (6.4,4).]

Another stability criterion referred to in Chap. 2 is the derivative dCmldCL(2.3,8). It was pointed out there that this derivative can only be said to exist if enoughconstraints are imposed on the independent variables a, V, 8e, q, etc., on which Cm

and CL separately depend. Such a situation results if we postulate that the vehicle isin rectilinear motion (q == 0) at constant elevator angle and throttle setting, with L =W, but with varying speed and angle of attack. Such a condition cannot, of course,actually occur in flight because the pitching moment could be zero at only one speed,but it can readily be simulated in a wind tunnel where the model is restrained by abalance. With the above stipulations, Cm and CL reduce to functions of the two vari-ables u and a, and incremental changes from a reference state ( )0 are given by

dCL = CLa da + CLu dfldCm = Cma da + Cmu dfl (6.4,5)

The required derivative is then

dflCma + Cmud;;

dflCLa + CLu da

provided dulda exists. This is guaranteed by the remaining condition imposed, thatis, L = W (implying aT == 0). For then we have

W = CL(a, u)tpV2S = const

(6.4,6)

from which we readily derive

(CLada + CLudfl)tPu't;> + CLQPUoSdu = 0 (6.4,7)

From (6.4,7)

ordfl CLa-=-da CLu + 2CLo

(6.4,8)

After substituting (6.4,8) into (6.4,6) and simplifying we get

ac; I 1dC = 2C C [Cma(CLu + 2CLo) - CLaCmJL L=W Lo La

On comparing (6.4,9) with (6.4,4), with the same caveats as for the trim slope, we seethat the static stability criterion is

(6.4,9)

Page 187: Stability Control EtkinReid

6.5 Effect of Flight Condition on the Longitudinal Modes of a Subsonic Jet Transport 177

dCml-- <0dCL L=W

provided that dC,jdCL is calculated with the constraints aSe = asp = q = 0 and L ==W. [The quantity on the left side of (6.4,9) is sometimes referred to as speed stabilityin the USA, by contrast with "angle of attack" stability. In Great Britain, this termusually has a different meaning, as in Sec. 8.5.]

On using the definition of h, given in (2.4,26) we find from (6.4,9) that

dC I ( C)~ = l+~ (h-h)dCL L=W 2CLo s

(6.4,10)

(6.4,11)

that is, that it is proportional to the "stability margin," and when CL• <t 2CLo' is equalto it.

6.5 Effect of Flight Condition on the LongitudinalModes of a Subsonic Jet Transport

In Sec. 6.2 we gave the representative characteristic modes of a subsonic jet airplanefor a single set of parameters. It is of considerable interest to enquire into how thesecharacteristics are affected by changes in the major flight variables-speed, altitude,angle of climb, and stability margin. It is also of interest to look into the effect of thevertical density gradient in the atmosphere. In this section and by means of exerciseswe examine some of these effects.

EFFECT OF SPEED AND ALTITUDE

The data in Heffley and Jewel (1972) for the example airplane include several com-parable cases, all having the same geometry, static margin, and gross weight. Thereare two speeds at sea-level, and three each at 20,000- and 40,OOO-ftaltitudes. Themodal periods and damping for these eight cases are displayed in Fig. 6.5. (Sincethere are so few points, the shapes of the curves are conjectural!) It is an understate-ment to say that there is no simple pattern to these data. It can be said, however, thatthe phugoid period increases with speed, as predicted by the Lanchester theory, anddecreases with altitude at fixed Mach number. The short-period does the opposite, de-creasing with speed and increasing with altitude.

The most striking feature of the data is the sudden and large increase in thephugoid period at high Mach number at the two higher altitudes. This phenomenon isa result of a loss of true static stability at these Mach numbers brought about by anegative value of Cm.' which has the effect of reducing E in (6.4,2). This happens be-cause this large aircraft is necessarily quite flexible, and because at these Mach num-bers it is entering the transonic regime, where air compressibility leads to substantialalterations in the aerodynamic pressure distribution. To show that Cm. is the reasonfor the behavior of the graphs, we vary it over a large range for the flight condition M= 0.8 and 20,000-ft altitude. Figure 6.6 shows the result and substantiates the impor-tant role of this derivative. In fact, from (6.4,4) we calculate that E = 0 when Cm. =-0.0968, a value only 4% more negative than that of the example at the given flight

Page 188: Stability Control EtkinReid

178 Chapter 6. Stability of Uncontrolled Motion

700,--------------- --.,

600

500

400

300

200

100

---- Period (s)

20.000ft

Mach number

(b) Short-period mode

Figure 6.5 Variation oflongitudinal modes with speed and altitude. (a) Phugoid mode. (b) Short-period mode.

----- Nha1tx20

40.000ft

~-~-----~~ '~----"~ - '20.0 0,,,~,OOOo L- __ ----L --L ...L .••~ __ ____=:'~ _

0.4 0.5 0.6 0.7 0.8 0.9

Mach number

(a) Phugoid mode

10...------------------------.

8 «iooo n

20,000ft

o~~ - , ..•- "-._--".6 / 40,000

4----- ---, " "

" 20,000------ ----02

---- Period (s)

----- Nha1tx 20

O'-- __ ..,-L-__ ---I. .L-__ ---L. __ ----J ---J

0.4 0.6

Page 189: Stability Control EtkinReid

6.5 Effect of Flight Condition on the Longitudi1Ul1 Modes of a Subsonic Jet Transport 179

700.-----------------------,

• B 747 example600

500

400

VIe-,"

300

200

100

0-0.1

IIIII,,\\\\\\\,,

"-

-------~~~~-- ---~~~~~~~~~~~~-- - =~::::::-:::~-----------Approx. (6.3.107--

(a) Period

5.----------------------....,

4

3

2-------

------'Approx. (6.3.10)

0L- ---L. ..I..- ---l ...L- __-0.1 -0.05 o 0.05 0.1

(b) Damping

Figure 6.6 Effect of Cm" on the phugoid mode. (a) Period. (b) Damping.

Page 190: Stability Control EtkinReid

180 Chapter 6. Stability of UncontroUedMotion

condition. Thus at this point the airplane would be very close to the static stabilityboundary, at which the period would go asymptotically to 00.

EFFECT OF VERTICAL DENSITY GRADIENT

We might expect on physical grounds that the vertical gradient in atmospheric den-sity would have an effect on the phugoid mode. For when the airplane is at the bot-tom of a cycle and moving fastest it is also in air of greater density and hence wouldexperience an additional increase in lift. It turns out that this effect is appreciable inmagnitude. We shall therefore do two things: (1) show how to include this effect inthe general equations of motion, and (2) derive a representative order of magnitude ofthe change in the phugoid period.

The modification to (4.9,18) consists of moving the ~ZE equation into the matrixequation and adding some appropriate derivatives to the aerodynamic forces (4.9,17).Since the only possible steady reference state in a vertically stratified atmosphere ishorizontal flight, we take (Jo = O.If A denotes the original system matrix, the result is

m W

1 [ MwZz] q- M + AnIy z m - Zw L.lU

ZEo--------------,-----------------o 1 0 -Uo: 0

In (6.5,1) there are three new derivatives with respect to ZE' Consider Zz first:

Z = CztpV2S

A ~u

(6.5,1)

az az apaZE = ap aZE

= iV2S [c + p acz] apz ap aZE

It is reasonable to neglect the variation of C, with p, and the density varies exponen-tially with height.' so that

(6.5,2)

apand - = «p (6.5,3)aZEwhere K is constant over a sufficient range of altitude for a linear analysis. It followsthat

2Exactly, in an isothermal atmosphere of uniform composition; approximately, in the real atmos-phere.

Page 191: Stability Control EtkinReid

6.5 Effect of Flight Condition on the Longitudinal Modes of a Subsonic Jet Transport 181

oZ- = ~KpV2SCzOZE

z, = oZ I = doOZE 0

(6.5,4)

Similarly,

X, = KXoMz = KMo

From (4.9,6) we get the reference values, leading to

Zz = <mg«Xz = M, = 0

The result is a rather simple elaboration of the original matrix equation. To get an es-timate of the order of magnitude of the density gradient effect, it is convenient to re-turn to the Lanchester approximation to the phugoid, and modify it to suit.

In Sec. 6.3 we saw that with this approximation, there is a vertical "spring stiff-ness" k given by (6.3,3) that governs the period. When the density varies there is asecond "stiffness" k' resulting from the fact that the increased density when the vehi-cle is below its reference altitude increases the lift, and vice versa. This incrementallift associated with a density change is

AL = CLiv2s dp

(6.5,5)

so that

Using (6.5,3) we get

k' = KW (6.5,6)

Thus we find that k' is approximately constant, whereas k from (6.3,3) depends onCwoP' which varies as y-2 for constant weight. The density gradient therefore has itsgreatest relative effect at high speed. The correction factor for the period, whichvaries inversely as the square root of the stiffness, is

(/k )112 1

F = k + k' = (1 + k' /k)ll2 (6.5,7)

so that the period, when there is a density gradient, is T' = FT. With the given valuesof k and k' this becomes

1

(1 + ~:%rl2in which the principal variable is seen to be the speed. Using a representative valuefor K of 4.2 X 10-5 (6.5,8) gives a reduction in the phugoid period of 18% for the ex-ample airplane at 774 fps. This is seen to be a very substantial effect. If the full sys-

F= (6.5,8)

Page 192: Stability Control EtkinReid

182 Chapter 6. Stability of Uru:ontrolled Motion

a '-----"'---- ---J... ...L- ----l ...J aa 0.05 0.1 0.15 0.2 0.25

100

'"~'

50

Exact 5

T

---------------_----- (6.3,10) 4

2-------------------_---- (6.3,10)Nha1f--

Static margin, Kn

Figure 6.7 Variation of period and damping of phugoid mode with static margin.

tern model (6.5,1) is used, comparable effects can be found on the damping of thephugoid as well.

EFFECTOFCGLOCATIONIt was indicated in Chap. 2 that the single most important aerodynamic characteristicfor longitudinal stability is the pitch stiffness Cma, and that it varies strongly with theCG position, that is,

Cma = CLa(h - hn)

where the static margin is K; = b; - h. The effect of this parameter is demonstratedby using (4.9,18) with variable Kn• The results, with all the other numerical data iden-tical with that in Sec. 6.2, are shown in Figs. 6.7-6.9. Figure 6.7 shows that thephugoid period and damping vary rapidly at low static margin and that the approxi-mation (6.3,10) is useful mainly at large Kn. Figure 6.8 shows the variation of the

0.3

100

0.2T

tJl '".•~' :.t

500.1

Exact and (6.3,15)

a '-- -'---- ----'- --'- '-- --J 0o 0.05 0.1 0.15 0.2 0.25

Static margin, K;Figure 6.8 Variation of period and damping of short-period mode with static margin.

Page 193: Stability Control EtkinReid

.15

.12

.09

.06

.03~ -0.005 -0.007~ 0S"

-.03

-.06

-.09

-.12

-.15-.15

.50

.40••

.30

.20

~ .10~

-.10

-.20

-.30

-.40

-.50-.80

.15

.12

.09

.06

.03~~ 0S"

-.03

-.06

-.09

-.12

-.15-.03

-Root locus...."..

0.04-

Oscillation branch -"? Static margin/(h,,-h)

- 0.02

~-0.02 0 0.0075 0 -0.02

CSubsidence /

A B~ -0.07

branchf- 0.02

~0.04

f-

I I I I I I I

o

-.60 o

••

-.20

(a)

Phugoidbranch

-0.07 -0.06

o -0.005

.03

(e)

Figure 6.9 (a) Locus of short-period roots, varying static margin. (b) Locus of phugoid roots,varying static margin. (c) Locus of phugoid roots, varying static margin, Mu = O.

-0.06

o

n, S-1

(b)

~_--_-:-0.3

------·-0.3

183

Page 194: Stability Control EtkinReid

184 Chapter 6. Stability of Uncontrolled Motion

short-period roots. These, too, vary strongly with pitch stiffness, the mode becomingnonoscillatory at a static margin near zero. The approximation (6.3,15) is excellentover the whole oscillatory range.

Important additional insight into these modes is obtained by examining the rootloci obtained by varying the static margin. These are shown in Fig. 6.9. Figure 6.9ashows that the damping, n, of the short-period mode remains virtually constant withdecreasing K; while the frequency, cu, decreases to zero at point A where the locussplits into two real roots, branches AB and AC of the locus. These of course representnonperiodic modes or subsidences. Figure 6.9b displays the much more complex be-havior of the phugoid. With reducing static margin (rearward movement of the CG)this mode becomes unstable at point D. At (totally unrealistic!) negative static marginbeyond -0.1, a new stable oscillation has reappeared. However, it is accompanied bya catastrophic positive real root far to the right.

The importance of the Mu derivative was shown earlier. It is again displayed inFig. 6.9c, which repeats the locus of the phugoid roots with Mu set equal to zero. Thecorresponding locus for the short-period roots is almost identical to Fig. 6.9a. Thepattern with M; = 0 would be more representative of a rigid airplane at low Machnumber. It shows a stable phugoid at all static margins as K; is decreased until itsplits at point D into a pair of real roots. The left branch from D then interacts withthe branch AB of the short-period locus to generate a new stable oscillation while theright branch crosses the axis to give an unstable divergence at negative static margin.

6.6 Longitudinal Characteristics ofa STOLAirplaneThe curves of Fig. 6.5 show that the characteristic modes of an airplane vary withspeed, that is with the equilibrium weight coefficient Cwo' In particular, the two char-acteristic periods begin to approach one another as Cwo becomes large. It is of inter-est to explore this range more fully by considering a STOL airplane, operating in the"powered-lift" region for which Cwo may be much larger. To this end the data givenin Margason et al., (1966) has been used to obtain a representative set of coefficientsfor 2.0 -s Cwo $ 5.0. The flight condition assumed is horizontal steady flight, so thatCxo = O. (The particular data used for the reference was that for the aircraft with alarge tail in the high position, it = 0, and 5f = 45°.) From the given curves, and fromcross-plots of the coefficients Cv CD' and Cm vs. CT at constant a, the data in Table6.5 were derived for the equilibrium condition. Smooth curves were used for interpo-lation. Since this is not a tilt-wing airplane, aT is not large in the cases considered,and has been assumed to be zero.

Since aeroelastic and compressibility effects are negligible at the low speeds ofSTOL flight the required speed derivatives are given by (see Table 5.1)

aCL sc;CZu = - CTu aCT; Cmu = CTu aCT

••

For a propeller-driven airplane, the value of CTuis given by (5.3,6), and an examina-tion of the data for a typical constant-speed propeller at low speed? showed thataT/au is very small. Hence we have used CTu = -2CTo in this example.

3The De Havilland Buffalo airplane.

Page 195: Stability Control EtkinReid

6.6 Longitudinal Cluuacteristics of a STOL Airplane 185

..

Table 6.5Basic Data for STOL Airplane

CLa CD,,} hx.* «; deD demCwo CTo deT sc; sc;2.0 0.53 5.75 1.19 0.500 0.705 0.285 -0.0902.5 0.72 6.20 1.80 0.475 0.790 0.328 -0.0703.0 0.90 6.65 2.41 0.450 0.875 0.370 -0.0503.5 1.09 7.10 3.02 0.424 0.955 0.411 -0.0304.0 1.28 7.55 3.63 0.398 1.025 0.450 -0.0104.5 1.46 8.00 4.24 0.371 1.097 0.488 +0.0105.0 1.65 8.45 4.85 0.346 1.165 0.525 +0.030

aCDtCD •• = aC

LCL.,

:t:hn= 0.30 - C••••ICL.,

Using the formulae of Table 5.1, the following estimates were made of the q anda derivatives:

CZq = -14, Crnq = -17.9, Cza = -5.5,

Finally the following inertial and geometric characteristics were assumed:

W = 40,0001b (177,920 N), S = 1000 fe (92.9 nr'),

A = 5.42, C = 13.60 ft (4. 145m),JL = 76.8, t,= 385, h = 0.30

With the above data, the coefficients of the system matrix were evaluated, and itseigenvalues and eigenvectors calculated. The main results are shown on Figs.6.10-6.13. Figures 6.10 and 6.11 show the loci of the roots as Cwo varies between 2and 5. The effect of Cwo is seen to be large on both modes, the short-period mode be-coming nonoscillatory at a value of Cwo somewhat greater than 3.5, and the damping

100wt*

54321o--~-=-.....•;.:c-~t--!:,o--~""'o-:!-~!:-"""""::--+----IO-1-2-3-4-5

Figure 6.10 Root locus-short-period mode, STOL airplane.

Page 196: Stability Control EtkinReid

186 Chapter 6. Stability of Uncontrolled Motion

--:::'::---:'::---=':--::'-:---="=------:::'=---="'=--I-;;o~---~ 100nt"

-0.5

-1.0

-1.5-2.0-2.5

100 wt'

4.03.5~.o 2.5

2.0Cwo

oCwo/..2.0--.......--Jl·5

4.5 4.03~0

5.0

Figure 6.11 Root locus-phugoid mode, STOL airplane.

of the phugoid increasing rapidly at the same time. Figure 6.12 shows the two peri-ods, and that they actually cross over at Cwo == 3.4. The concept of the phugoid as a"long" period oscillation is evidently not applicable in this situation! The approxima-tions to the phugoid and the pitching mode are also shown for comparison. It is seenthat they give the two periods quite well, and that (6.3,15) also depicts quite accu-rately the damping of the pitching oscillation and of the two nonperiodic modes intowhich it degenerates at high Cwo' The phugoid damping, however, is not at all wellpredicted by the approximate solution. Figure 6.13 shows that the modes are all heav-ily damped over the whole range of Cwo'

..

\

40

30

20

10

°1~-----:2!:------:3:-------4';-----1\5

CwoFigure 6.12 Periods of oscillatory modes, STOL airplane.

Page 197: Stability Control EtkinReid

6.7 Lateral Modes of a Jet Transport 187

5

Exact

10

.•...•..<, )

PhlllOlcf•••••••~.3.1 0......--.•.•.•..•...... -..Nonoscillitory ------modes (x 10)

25

20

••,,; 15J.

o":::-----""'*------:::I:------:'-:::-------:lU W ~ ~ ~Cwo

Figure 6.13 Time to damp of modes, STOL airplane.

6.7 Lateral Modes of a Jet TransportWe use the same airplane and flight condition as for the longitudinal modes in Sec.6.2, and calculate the lateral modes. The nondimensional and dimensional derivativesare given in Tables 6.6 and 6.7. Using these, the system matrix of (4.9,19) is found tobe (note that the state vector is [v p r <t>Y.):

[

-0.0558 0A = -0.003865 -0.4342

0.001086 -0.006112o 1

-7740.4136

-0.1458o TJ

This yields the characteristic equation

A4 + 0.6358A3 + 0.9388A2 + 0.5114A + 0.003682 = 0 (6.7,2)

The stability criteria are

E = 0.003682 > 0R = 0.04223 > 0

so there are no unstable modes.

Table 6.6Nondimensional Derivatives-B747 Airplane

-0.8771oo

-0.2797-0.3295

0.304

0.1946-0.04073-0.2737

Page 198: Stability Control EtkinReid

188 Chapter 6. Stability of UncontroUed Motion

Tabk6.7Dimensional Derivatives-B747 Airplane

Y(lb) L(ft'lb) N(ft'lb)

v(ftls) -1.103 x 103 -6.885 X Ht 4.790 X Htp(radls) 0 -7.934 X 106 -9.809 X HPr(radls) 0 7.321 X 106 -6.590 X 106

YeN) L(m·N) N(m'N)

v(rnls) -1.610 X 104 -3.062 X HP 2.131 X HPp(radls) 0 -1.076 X 107 -1.330 X 106

r(radls) 0 9.925 X 106 -8.934 X 106 ..

EIGENVALUES

The roots of (6.7,2) areMode 1 (Spiral mode):

Mode 2 (Rolling convergence):

Mode 3 (Lateral oscillation or Dutch Roll):

Al = -0.0072973A2 = -0.56248

A3•4 = -0.033011 ± 0.94655i

Table 6.8 shows the characteristic times of these modes. We see that two of them areconvergences, one very rapid, one very slow, and that one is a lightly damped oscilla-tion with a period similar to that of the longitudinal short-period mode.

"EIGENVECTORS

The eigenvectors corresponding to the above eigenvalues are given in Table 6.9. Inaddition to the basic 4 state variables, Table 6.9 contains two extra rows that show thevalues of the two state variables l/J and YE (see Exercise 6.2).

MODE 1: THE SPIRAL MODE

From Table 6.9, we find the ratios of the angle variables in the spiral mode to be

f3:<fJ:l/J = -0.00119:-0.177:1

so that the motion is seen to consist mainly of yawing at nearly zero sideslip withsome rolling. This is, of course, the condition for a truly banked turn, and this mode

Table 6.8Characteristic Times-Lateral Modes

Mode Name Period (s)

12

SpiralRolling

convergenceLateral oscillation

(Dutch Roll)

951.23

3 6.64 21 3.16

Page 199: Stability Control EtkinReid

6.7 Lateral Modes of a Jet Transport 189

Table 6.9Eigenvectors (polar form)

Spiral Rolling convergence Dutch Roll

Magnitude Phase Magnitude Phase Magnitude Phase

f3=D 0.00119 180° 0.0198 180° 0.33 -28.1°P 1.63 X 10-4 0° 0.0712 180° 0.12 92.0°r 9.20 X 10-4 180° 0.0040 0° 0.037 -112.3°cP 0.177 180° 1.0 0° 1.0 0°I/J 1.0 0° 0.0562 180° 0.31 155.7°YE 7.772 X 1<>3 180° 7.65 0° 1.69 -165.8°

uot*

can be thought of as a variable-radius turn. The aerodynamically important variablesare

{3:p:r = 1:-0.137:0.773

and the largest of these, {3,has already been seen to be negligibly small for moderatevalues of t/J and 1/1. The aerodynamic forces in this mode are therefore very small, andit may be termed a "weak" mode. This is consistent with its long time constant.

The flight path in the spiral mode can readily be constructed for any given initialyaw angle from the eigenvector. For example, with an initial 1/1 of 20° (0.35 rad), wehave from Table 6.9

1/1 = 0.35eA1t

v = 774( -0.OO119)0.35eA1t fpswhere A1= -0.0072973 S-1

From (4.9,19) and the above it follows that (for 00 = 0)

YE = -3707ge-o.OO72973t ft

XE = 774 tft

Figure 6.14 shows the path-it is seen to be a long, smooth return to the referenceflight path, corresponding to YE = O. When the spiral mode is unstable as is fre-quently the case, 1/1, t/J, and YE are all of the same sign, and of course all increase withtime instead of decreasing, as shown in the figure.

MODE 2: THE ROLLING CONVERGENCE

The ratios of the angle variables in this mode are, from Table 6.9,

{3:t/J: 1/1 = -0.0198:1:-0.0562

The mode is evidently one of almost pure rotation around the x axis, and hence itsname. The variables that are significant for aerodynamic forces are ({3,p, r) and theyare in the ratios

{3:p:r = 0.278:1:-0.0561

Page 200: Stability Control EtkinReid

190 Chapter 6. Stability of Uncontrolled Motion

--+--==-=,.......,,!-....,..,,..,,..,,.,:--+---- YEo ItI. 37,079 gl. 37.079 ••IFigure 6.14 Flight path in spiral mode.

so that the largest rolling moment in this mode of motion is C, p, and the f contribu-p

tions are negligible by comparison.

MODE 3: THE LATERAL OSCILLATION (DUTCH ROLL)

The vector diagram for this mode is shown in Fig. 6.15. It is seen that the three anglevariables {3,¢, l/J are of the same order of magnitude, that f is an order smaller, and

1m

tfl = 1.0-------+-::::::sl::l~========> ~Re

lJF= 0.31

f3 = ~ = 0.33

~ = 0.037(not visible)

Figure 6.15 Vector diagram oflateral oscillation. Cwo = 0.57. Altitude =i= 40,000 ft.

Page 201: Stability Control EtkinReid

6.7 Lateral Modes of a Jet Transport 191

that {3and I/J are almost equal and opposite. It follows from (4.9,19) that YE is nearlyzero. In dimensional terms, when 1</>1 = 20°, ~EI== 8 ft, whereas the wavelength ofthe oscillation is about 5000 ft. The vehicle mass center is seen to follow a nearlyrectilinear path in this mode, the motion consisting mainly of yawing and rolling, thelatter lagging the former by about 160° in phase.

EFFECT OF SPEED AND ALTITUDE

Even for the "basic" case of a rigid airplane at low Mach number, the variation of thelateral modes with speed and altitude may not be simple. This is because some of thelateral stability derivatives are dependent on the lift coefficient in complex ways. Thatis especially true of airplanes with swept wings and low aspect ratio for which C1p in-creases markedly with CL' These effects will appear most strongly at low speed andhigh altitude, both of which require high CL• (Note that in the B747 example at M =0.8 and 40,000 ft, CL = 0.654, which is quite large for cruising flight.) For a rigidswept-wing airplane at low Mach number the period of the Dutch Roll mode wouldbe expected first to increase and then to decrease as the airplane speed increases. Thedamping of this mode would be expected to be weak at low speed and to increase athigher speeds. The rolling convergence is well damped at all speeds, but the dampingwould normally increase with speed. The spiral mode is frequently unstable oversome portion of the speed/altitude flight envelope, depending on the interplay of thederivatives that appear in (6.8,6). The characteristic times of this mode are, however,usually so long that the instability does not degrade the handling qualities unduly.

The effect of increasing altitude at fixed CL is primarily an increase in the damp-ing time constants of all the modes. The period of the Dutch Roll is not much af-fected.

When substantial aeroe1astic and compressibility effects are added to the alreadycomplex behavior of the lateral modes, the result is an even more irregular pattern ofmodal characteristics. The data of (Heffley and Jewel, 1972) for the B747, repro-duced in Table 6.10 show this. (Note that a negative thalf implies an unstable mode.)At the two lower altitudes, with relatively low values of CL> the modes are seen to be-have in a fairly regular way (see Fig. 6.16). However, at 40,000 ft and high Mach

Table 6.10Variation of Lateral Modes with Speed and Altitude

Spiral Rolling Lateral oscillationmode convergence (Dutch Roll)

Altitude, Mach thalf thalf Period Nha/fIt No. (s) (s) (s) (cycles)

0 0.45 35.7 0.56 5.98 0.870 0.65 34.1 0.44 4.54 0.7120,000 0.5 76.7 0.93 7.3 1.5820,000 0.65 64.2 0.76 5.89 1.3320,000 0.8 67.3 0.85 4.82 1.1240,000 0.7 -296 1.5 7.99 1.9340,000 0.8 94.9 1.23 6.64 3.1540,000 0.9 -89.2 1.45 6.19 1.18

Page 202: Stability Control EtkinReid

192 Chapter 6. Stability of Uncontrolled Motion

8r--------------------~7

6

5

4

3

2

Altitude20,000 ft

- 20,000ft

------0OL-_-----JL-_-----J __ -----J__ ----l__ ----l__ ----l__ ----I0.3 0.4 0.6 0.90.7 0.80.5

Mach number

(a) Lateral oscillation

1.4

0L-__ L-__ L-__ L-__ L-__ L-__ L-__0.3 0.4 0.5 0.6 0.7 0.8 0.9

Mach number(b) Rolling convergence

Figure 6.16 Variation oflateral modes with speed and altitude. (a) Lateral oscillation. (b) Rollingconvergence. (c) Spiral mode.

1.2

1.0

Ul. 0.8'"'.J

0.6

0.4

0.2

Altitude20,000 ft-------0

Page 203: Stability Control EtkinReid

6.8 Approximate Equations for the Lateral Modes 193

100r--------------------,

80

"- Altitude~20,000ft

60

40 - _0

20

00.3 0.4 0.5 0.6 0.7 0.8

Mach number

(e) Spiral mode

Figure 6.16 (Continued)

number the lateral behavior is quite irregular, especially the variation of the dampingof all the modes with M. The spiral mode is seen to be unstable (albeit with long timeconstant) at both M = 0.7 and 0.9 but stable at M = 0.8. This behavior is primarilythe result of the complex variation of C1p with CL and M in this region.

6.8 Approximate Equations for the Lateral ModesAs with the longitudinal modes we should like if possible to have useful analyticalapproximations to the lateral characteristics. We find that there are reasonable ap-proximations to all three modes, but the application of all such approximations mustbe made with caution. Their accuracy can really be verified only a posteriori, bycomparison with exact solutions. They can only be used with confidence in situationssimilar to those in which they have previously been found to work well.

SPIRAL MODE

Comparison of the eigenvalues in Sec. 6.7 shows that A for the spiral mode is two or-ders of magnitude smaller than the next larger one. This suggests that a good approx-imation to this root may be obtained by keeping only the two lowest-order terms inthe characteristic equation, that is,

or

DA + E==OAs == -E/D

(6.8,1)

where As denotes the real root for the spiral mode. Before deriving expressions for Dand E, we rewrite the matrix of (4.9,19) in a more compact notation for convenience,including the approximation Yp = O.

Page 204: Stability Control EtkinReid

194 Chapter 6. Stability of Uncontrolled Motion

A = [~: ~p ~: g c~s lJo]oN s oN 0 (6.8,2)

v p ,

o 1 tanlJo 0

The meanings of the symbols in (6.8,2) are obtained by comparison with (4.9,19), forexample

L:£ = --.!!... + I'zxN,v I~ v

and in the special case when the stability axes are also principal axes, Ivc = 0 and

Lv:£=-v I,

With the notation of (6.8,2), expanding det (A - AI) yields

E = g[(:£voN, - :£,oNv) cos lJo+ (:£poNv - :£voNp) sin 60] (a)

D = -g(:£v cos lJo+ oNv sin lJo) + ayv(:£,oNp - :£pX,) (6.8,3)

+ ay.(:£pXv - :£voNp) (b)

When the orders of the various terms in D are compared, it is found that the secondterm can be neglected entirely and Yr can be neglected in ayr' The approximation thatthen results is

D = -g(:£v cos 60 + oNv sin (0) + uo(:£voNp - :£pXv) (6.8,4)

The result obtained from (6.8,1), (6.8,3a), and (6.8,4) for the jet transport example ofSec. 6.7 is As = -0.00725, less than 1% different from the correct value. Equation(6.8,1) is seen to give a good approximation in this case.

It will be recalled that the coefficient E has special significance with respect tostatic stability (see Sec. 6.1). We note here that in consequence of (6.8,1) the spiralmode may exhibit exponential growth, and that the criterion for static lateral stabilityis

(:£voN, - :£,.Nv) cos 60 + (:£p.Nv - :£v.Np) sin 60> 0 (6.8,5)

On substituting the expanded expressions for:£v and so forth, (6.8,5) reduces to

(CZ,Pnr - CZrCnp) cos 60 + (CZpCnp - CZpCnp) sin 60> 0 (6.8,6)

Since some of the derivatives in (6.8,6) depend on CLo, the static stability will varywith flight speed. It is not at all unusual for the spiral mode to be unstable over someportion of the flight envelope (see Table 6.10 and Exercise 6.3).

ROLLING MODE

It was observed in Sec. 6.7 that the rolling convergence is a motion of almost a singledegree of freedom, rotation about the z-axis, This suggests that it can be approxi-mated with the equation obtained from (4.9,19) by putting v = r = 0, and consider-ing only the second row, that is,

Page 205: Stability Control EtkinReid

6.8 Approximate Equations for the lAteral Modes 195

p = ;£pP

which gives the approximate eigenvalue

AR == ;£p = L/ I~ + I~xNp

(6.8,7)

(6.8,8)

The result obtained from (6.8,8) for the B747 example is AR = -0.434, 23% smallerthan the true value -0.562. This approximation is quite rough.

An alternative approximation has been given by McRuer et al. (1973). This ap-proximation leads to a second-order system, the two roots of which are approxima-tions to the roll and spiral modes. In some cases the roots may be complex, corre-sponding to a "lateral phugoid"-a long-period lateral oscillation. The approximationcorresponds to the physical assumption that the side-force due to gravity producesthe same yaw rate r that would exist with (3 = O. Additionally Yp and Yr are ne-glected. With no approximation to the rolling and yawing moment equations the sys-tem that results for horizontal flight is

0= -uor + gc/JP = ;£vv + ;£pP + ;£7r = Nvv + Npp + Nrr¢=P

(a)(b)(c) (6.8,9)

(d)

The procedure used to get the characteristic equation of (6.8,9) is the same as thatused previously for the phugoid approximation in Sec. 6.3. The result is

goo-A

=0 (6.8,10)

which expands to

CA2 +DA +E= 0

C = uoNvD = uo(;£vNp - ;£~v) - g;£vE = g(;£vNr - ;£rNv)

The result of applying (6.8,11) to the B747 example is

As= -0.00734 and AR = -0.597

where(a)(b)

(c)

(d)

(6.8,11)

These are within about 1% and 6% of the true values, respectively, so this is seen tobe a good approximation for both modes, certainly much better than (6.8,7) for therolling mode.

DUTCH ROLL MODE

A physical model that gives an approximation to the lateral oscillation is a "flat"yawing/sideslipping motion in which rolling is suppressed. The corresponding equa-tions are obtained from (4.9,19) by setting P == c/J == 0 and dropping the second(rolling moment) equation. The term in Yr is also neglected in the first equation. The

Page 206: Stability Control EtkinReid

196 Chapter 6. Stability of Uncontrolled Motion

result is

v = CfYvv - uori = ./fvv + ./frr

The corresponding characteristic equation is readily found to be

11.1 - (GYv + ./fr)A + (GYv./fr + uo./fv) = 0 (6.8,12)

The result obtained from (6.8,12) for our example is AVR = -0.1008 ::t 0.9157i, or

T = 6.86 secNhalf = 1.0

The approximation for the period is seen to be useful (an error of about 3%) but thedamping is very much overestimated.

There is another approximation available for the damping in this mode that maygive a better answer. It follows from the fact that the coefficient of the next-to-highestpower of A in the characteristic equation is the "sum of the dampings" (see Exercise6.4). Thus it follows from the complete system matrix of (6.8,2) that

2nVR + AR + As= GYv + Xp + ./fror nVR = HGYv + Xp + ./fr - (AR + As)} (6.8,13)

But the approximation (6.8,11) for the roll and spiral modes gives precisely

DAR + As = --

C

On using (6.8,11) we get the expression

nVR = ~ [GYv + ./fr + ~: (./fp - ~)] (6.8,14)

which is to be compared with ~(GYv + ./fr) given by (6.8,12). The damping obtainedfrom (6.8,14) is nVR = -0.0159, better than that obtained from (6.8,12) but still quitefar from the true value of -0.0330. The simple average of the two preceeding ap-proximations for the Dutch Roll damping has also been used. In this instance it givesnVR = -0.0584, which although better is still 77% off the true value.

This example of an attempt to get an approximation to the Dutch Roll dampingillustrates the difficulty of doing so. Although the approximation tends to be better atlow values of CL> nevertheless it is clear that it must be used with caution, and thatonly the full system matrix can be relied on to give the correct answer.

\ ..•.

6.9 Effects of WindIn all the preceding examples, the atmosphere has been assumed to be at rest or tohave a velocity uniform in space and constant in time. Since this is the exceptionalrather than the usual case, it is necessary to examine the effects of nonuniform andunsteady motion of the atmosphere on the behavior of flight vehicles. The principaleffects are those associated with atmospheric turbulence, and these are treated atsome length in Etkin (1972, 1981). However, quite apart from turbulence, the wind

Page 207: Stability Control EtkinReid

6.9 Effects of Wind 197

may have a mean structure which is not uniform in space, that is, there can be spatialgradients in the time-averaged velocity. The examples of most concern are down-bursts and the boundary layer next to the ground produced by the wind blowing overit. Downbursts are vertical outflows from low level clouds that impinge on theground, somewhat in the manner of a circular jet, and spread horizontally. The result-ing wind field has strong gradients, both horizontal and vertical. A number of air-plane accidents have been attributed to this phenomenon.

In order to introduce wind into the analytical model, we must make any alter-ations that may be needed, because of the presence of the wind, to the aerodynamicforces and moments. In the trivial case when the wind is uniform and steady, nochange is necessary to the representation of aerodynamic forces from that used be-fore. However, turbulence and wind gradients may require such changes.

Since the linear model that was developed in Chap. 4 is based on small distur-bances from a steady reference condition, and since there is no such steady statewhen the aircraft is landing or taking off through a boundary layer or downburst, thelinear model is of limited use in these situations. For this kind of analysis, one mustuse the nonlinear equations (4.7,1)-(4.7,5) and introduce a model for the aerody-namic forces that embraces the whole range of speeds and attitudes that will occurthroughout the transient (Etkin, B. and Etkin, D. A. (1990». Such an analysis is be-yond the scope of this volume.

There is one relevant steady state, however, that can be investigated with the lin-ear model, and that is horizontal flight in the boundary layer. The planetary boundarylayer has characteristics quite similar to the classical flat-plate turbulent boundarylayer of aerodynamics. The vertical extent of this layer in strong winds dependsmainly on the roughness of the underlying terrain, but is usually many hundreds offeet. Figure 6.17 shows the power-law profiles associated with different roughnesses.These are all of the form

(6.9,1)

where, as indicated in the figure, h is height above the ground. The vertical gradientis then given by

dW-=nkhn-1

dh(6.9,2)

For example, for smooth terrain (n = 0.16), and for a wind of 50 fps at 50 ft altitude,the gradient would be dW/dh = 0.16 fps/ft.

By way of example, we shall analyze what effect the vertical wind gradient hason the longitudinal modes of the STOL airplane of Sec. 6.6 in low speed flight, whenwind effects can be expected to be largest. In order to generate the analytical model,we go back to the exact equations (4.7,1) et seq. The longitudinal equations, whenlinearized for small perturbations around a reference state of horizontal flight atspeed Uo and 00 = 0 are

ax - mgO = mAw

AZ = m(wE - ugq)AM=IiJiJ=q

z;' = -ugo + wE

(6.9,3)

-------"----"-----"-- -- "-- ""--- -"--_._-.

Page 208: Stability Control EtkinReid

200 Chapter 6. Stability of Uncontrolled Motion

Cnr• Note that for upwind flight this leads to an unstable roll "stiffness" C1</> > 0where none existed before (C1</> is negative for downwind flight).

Reasonable estimates of the major changes in the basic derivatives associatedwith I' can be made from available aerodynamic theories, but a complete account ofthese is not currently available, and to develop them here would take us too far afield.Instead we simply incorporate the additional terms given by (6.9,7) into the longitu-dinal equations of motion and note the extent of the changes they make in the charac-teristic modes previously calculated. The appropriate matrix for this case is obtainedfrom (4.9,18) by adding the two terms containing I', and is given by (6.9,8).

[XU (Xw ) ]A = - -- - I' 0 (- g + fuo)

-~-------~----~;------------------ (6.9,8)

In (6.9,8), A' is the 3 X 4 matrix consisting of the last three rows of the matrix of(4.9,18), with (Jo = O.

An example of the results for the STOL airplane is shown in Figs. 6.18 and 6.19.The numerical data used was the same as in Sec. 6.6, with wind gradient variablefrom -0.30 fps/ft (the headwind case) to +0.30 fps/ft (the tailwind case). The effectson both the phugoid and pitching modes are seen to be large. A strong headwind de-creases both the frequency and damping of the phugoid, and a strong tailwindchanges the real pair of pitching roots into a complex pair representing a pitching os-cillation of long period and heavy damping.

100 cut'

3.04W-;u;-+0.30

2.0

1.0-0.3

100 nt'0

-1.0

-2.0

-3.0

Figure 6.18 Effect of wind gradient on phugoid roots-STOL airplane. Cwo = 4.0.

Page 209: Stability Control EtkinReid

6.10 Exercises 201

100 lilt"

1.5

Figure 6.19 Effect of wind gradient on short-period roots-8TOL airplane. Cwo = 4.0.

6.10 Exercises6.1 Use (6.3,10) to calculate the approximate period T; = 27flwn of the phugoid oscilla-

tion. Assume Mu = O.

6.2 Derive the entries for'" and (yEluot*) in Table 6.9 from those for (j and f.

6.3 The stability derivatives of a general aviation airplane are given in Table 7.2. The air-plane weighs 2400 Ib (10,675 N) and has a wing area of 160 ft2 (14.9 nr'). The flightaltitude is sea level. Calculate and plot the spiral stability criterion E as a function ofspeed (0.15 < CL < 1.7) for values of 00 = -10°,0°, 10°.

6.4 The characteristic equation is of order N. Prove that the coefficient of >..N-l is the neg-ative of the sum of the real parts of all the roots, and hence is aptly termed "the sumof the dampings."

6.5 Find the critical climb angle for spiral stability of the jet transport of Sec. 6.7. [Hint:start with (6.8,6)]. Having regard to its expected influence on the stability derivatives,state the effect on spiral stability in horizontal flight of increasing the wing dihedralangle.

6.6 Carry out the transformation WB = LBEW E to get (6.9,5).

6.7 Using the stability derivatives given in Table 7.2 for a general aviation airplane, cal-culate the lateral modes in the absence of gravity. The relevant data are:

W = 2400 Ib (10,675 N) I; = 170 slug-ff (230 kg-nr')S = 160 ft2 (14.9 nr') I,= 1,312 slug-fr' (1,778 kg-nr')

b = 30 ft (9.14 m) Izx = 0

Page 210: Stability Control EtkinReid

202 Chapter 6. Stability of Uncontrolled Motion

v = 150 knots (77.3 mls) 00 = 0altitude = sea level

Compare the results with those for gravity present.

6.8 Find the characteristic equation of the hovercraft of Exercise 4.10. Show that when itis statically unstable with both Me and L<I> positive it can be gyrostabilized (like aspinning top, i.e., solutions remain bounded) if H is large enough.

6.9 A conventional stable aircraft is on a steady descent to a landing on a shallow glideslope when the headwind suddenly vanishes. What initial condition problem de-scribes the subsequent motion? Describe qualitatively, from your knowledge of longi-tudinal natural modes, what the subsequent flight path will be if the elevator andthrottle controls remain fixed at their prior positions.

6.10 Show that if we neglect all the Y force derivatives, then the small-disturbance equa-tions (for (Jo = 0) yield the following approximation for the lateral displacement:

tlyFi-t) = g I: I: c/>(T)dr dt

6.11 Assume that in the lateral oscillation of an airplane, the modal diagram shows thatIl/JI > IcPl and the rfI vector leads the c/J vector by an angle between 900 and 1800

• Thisis reasonably representative of flight at low Mach number. Describe the relative mo-tion of the two wing tips. To simplify the situation, neglect damping, the motion asso-ciated with sideslip and with the forward motion of the airplane, assume the anglesare "small," and interpret "wing tip" to mean a point on the y-axis.

Hold a model airplane by its wing tips and practise until you can execute a motion ofthe type you have deduced. Observe it in a mirror. (The name Dutch Roll was givento this motion because of a perceived resemblance to that of an ice skater, Hollandbeing noted for this sport.)

6.12 The theory for a stable airplane shows that if the controls are neutral it will fly in asteady state on a straight line at constant speed. However, the development in the textassumed that the airplane produced no lateral aerodynamic force or moments whenthe controls were set to neutral and the lateral state variables were zero. Real air-planes cannot achieve this owing to design and construction factors. You are asked inthis exercise to examine the steady states that are possible for such an airplane.The steady state here is defined to be one in which the linear and angular velocitycomponents (u, v, W, p, q, r), the aerodynamic forces and moments (X, Y, Z, L, M, N)and the gravity force components (Xg, Yg, Zg) are all constant.

(a) Show that in the steady state the two Euler angles 0 and cP are also constant.(b) Since 0 is constant, it can be set to zero by a suitable choice of body axes. For

this case show that p = 0 and the angular velocity vector Cd is vertical in FE'(c) Starting with (4.5,8) and (4.5,9) develop the lateral steady state equations. As-

sume that in the steady state q, r, cP and v are small quantities. Assume that thecontrols are neutral and thus (since p = 0 in the steady state)

Y = Ya + Yvv + Y,.r

L = La + Lvv + L,.r

N = N; + Nvv + N,.r

••

Page 211: Stability Control EtkinReid

6.11 Additional Symbols Introduced in Chapter 6 203

where ( )a stands for effects due to lateral asymmetries in the aircraft. Whatcondition must be met in order for a unique steady state to exist? How is this con-dition related to E of (6.8,3a) when the body axes are principal axes?

(d) Relate the results of (c) to the flight of a statically stable hand-launched glider.

6.11 Additional Symbols Introduced in Chapter 6

h altitude

;tv Lv!I~ + I'JVv

:£p L/I~ + I'JVp.. ;tr L,JI~ + I'JVr

n real part of eigenvalue

Xv NjI~ +I'zh

Xp Np/I~ + I'zxLp

Xr N,JI~ + I'zxLr

T period of oscillationCiYv Yv!mCiYr Y/m - Uo

r vertical wind gradient, -dW/dhK density gradient parameter (see 6.5,3)A eigenvalueAVR eigenvalue of lateral oscillationAR eigenvalue of rolling convergence

As eigenvalue of spiral mode, damping ratioW imaginary part of eigenvalue, circular frequencyWn undamped circular frequency

Page 212: Stability Control EtkinReid

CHAPTER 7

Response to Actuation of theControls-Open Loop

7.1 General RemarksIn this chapter we study how an airplane responds to actuation of the primary con-trols-elevator, aileron/spoiler, rudder, and throttle. These are of course not the onlycontrols that can be incorporated in the design of an airplane. Also used, less fre-quently, are vectored thrust and direct lift controL Closely related to the control-re-sponse problem is the response of the airplane to an in-flight change of configurationsuch as flap deflection, lowering the undercarriage, releasing stores or armaments,deploying dive brakes, or changing wing sweep. The analysis of the response of theairplane to any of these uses methods generally similar to those that are described inthe following. In the remainder of this chapter, it is assumed that there is no wind.

LONGITUDINAL CONTROL

The two principal quantities that need to be controlled in symmetric flight are thespeed and the flight-path angle, that is to say, the vehicle's velocity vector. To achievethis obviously entails the ability to apply control forces both parallel and perpendicu-lar to the flight path. The former is provided by thrust or drag control, and the latterby lift control via elevator deflection or wing flaps. It is evident from simple physicalreasoning (or from the equations of motion) that the main initial response to openingthe throttle (increasing the thrust) is a forward acceleration, i.e, control of speed. Themain initial response to elevator deflection is a rotation in pitch, with subsequentchange in angle of attack and lift, and hence a rate of change of flight-path direction.When the transients that follow such control actions have ultimately died away, thenew steady state that results can be found in the conventional way used in perfor-mance analysis. Figure 7.1 shows the basic relations. The steady speed Vat which theairplane flies is governed by the lift coefficient, which is in turn fixed by the elevatorangle-see Fig. 2.19. Hence a constant 5e implies a fixed V. The flight-path anglel' = 0 - ax at any given speed is determined, as shown in Fig. 7.1, by the thrust.Thus the ultimate result of moving the throttle at fixed elevator angle (when the thrustline passes through the CG) is a change in l' without change in speed. But we sawabove that the initial response to throttle is a change in speed-hence the short-termand long-term effects of this control are quite contrary. Likewise we saw that themain initial effect of moving the elevator is to rotate the vehicle and influence 1',whereas the ultimate effect at fixed throttle is to change both speed and 1'. The short-term and long-term effects of elevator motion are therefore also quite different. The_

204

Page 213: Stability Control EtkinReid

7.1 General Remarks 205

lv,h-consl

~----_...J...._------~VVmcl

Figure 7.1 Basic performance graph.

total picture of longitudinal control is clearly far from simple, and the transients thatconnect the initial and final responses require investigation. We shall see in the fol-lowing that these are dominated by the long-period, lightly damped phugoid oscilla-tion, and that the final steady state with step inputs is reached only after a long time.These matters are explored more fully in the following sections.

LATERAL CONTROLThe lateral controls (the aileron and rudder) on a conventional airplane have threeprincipal functions.

1. To provide trim in the presence of asymmetric thrust associated with powerplant failure.

2. To provide corrections for unwanted motions associated with atmospheric tur-bulence or other random events.

3. To provide for turning maneuvers-that is, rotation of the velocity vector in ahorizontal plane.

The first two of these purposes are served by having the controls generate aero-dynamic moments about the x and z axes-rolling and yawing moments. For the thirda force must be provided that has a component normal to V and in the horizontalplane. This is, of course, the component L sin cP of the lift when the airplane isbanked at angle cP. Thus the lateral controls (principally the aileron) produce turns asa secondary result of controlling cP.

Ordinarily, the long-term responses to deflection of the aileron and rudder arevery complicated, with all the lateral degrees of freedom being excited by each. Solu-tion of the complete nonlinear equations of motion is the only way to appreciatethese fully. Certain useful approximations of lower order are however available.

Page 214: Stability Control EtkinReid

206 Chapter 7. Response to Actuation of the Controls-Open Loop

THE CONTROL EQUATIONS

Whereas the study of stability that was the subject of Chap. 6 is generally sufficientlywell served by the linear model of small disturbances from a condition of steadyflight, the response of an airplane to control action or configuration change can in-volve very large changes in some important variables, especially bank angle, pitchangle, load factor, speed, and roll rate. Consequently nonlinear effects may be presentin any of the gravity, inertia, and aerodynamic terms. An accurate system model ca-pable of dealing with these large responses must therefore begin with the more exactequations (4.7,1)-(4.7,4). These would normally be reorganized into first-order statespace form for subsequent integration by a Runge-Kutta or other integration scheme.Equations (4.7,3d-f) and (4.7,4) are already in the required form. However (4.7,1)and (4.7,2) need to be rearranged. In particular, (4.7,2a and c) need to be solved si-multaneously for p and i,The functional form that results is as follows:

itE = f(vE, wE, q, r fJ, X') + Xc (a)m

yiJE= f(uE, wE, p, r, fJ, </>, Y') + -=- (b) (7.1,1)

mZwe = f(l/, vE, p, q, fJ, cP, Z') + -=- (c)m

Lp = f(p, q, r, L', N') + I; + I'zxNc (a)x

(7.1,2)Me

q = f(p, r,M') + I (b)y

Ni = f(p, q, r, L', N') + -7- + I'J.,c (c)t;tJ = f(q, r, cP) (a)

(7.1,3)tb = f(p, q, r, fJ, cP) (b)

In the preceding equations, the subscript c denotes the control forces and moments,and the prime on force and moment symbols denotes the remainder of the aerody-namic forces and moments. The solution of these equations would require that anaerodynamic submodel be constructed for each case to calculate the forces and mo-ments at each computing step from a knowledge of the state vector, the control vec-tor, the current configuration, and the wind field. To follow this course in extensowould take us beyond the scope of this text, so for the most part the treatments thatfollow are restricted to the responses of linear invariant systems, that is, ones de-scribed by (4.9,20) with A and B constant viz

x = Ax + Be (7.1,4)

Although we are thereby restricted to relatively small departures from the steadystate, these responses are nevertheless extremely useful and informative. Not only dothey reveal important dynamic features, but when used in the design and analysis ofautomatic flight control systems that are designed to maintain small disturbances theyare in fact quite appropriate.

'.

Page 215: Stability Control EtkinReid

7.2 Response of LinearHnvariont Systems 207

Table 7.1Dimensional Control Derivatives

x Z M

8. eX8JpU~S eZ8}apu~S em••!PU~Sc

s, eX8 !PU~S eZ8p!pU~S em.lpu~SC'P ,

y L N

e: eY8Qipu~S eI8J.pu~Sb en •.ipu~Sb

s. eY8Jpu~s el8Jpu~Sb en s,ipu~Sb

In the examples that follow, we use {Be' Bp} for the longitudinal controls, elevatorand throttle; and {8a, 8r} for the lateral controls, aileron, and rudder. The aerody-namic forces and moments are expressed just like the stability derivatives in terms ofsets of nondimensional and dimensional derivatives. The nondimensional set is thepartial derivatives of the six force and moment coefficients {Cx' Cy, Cz' Cl, Cm' Cn}with respect to the above control variables, such as Cza"= aC/iJ8e or ClBa = aC/a8a,

and so on. The dimensional derivatives are displayed in Table 7.1.The powerful and well-developed methods of modern control theory are directly

applicable to this restricted class of airplane control responses. Before proceeding tospecific applications, however, we first present a review of some of the highlights ofthe general theory. Readers who are well versed in this material may skip directly toSec. 7.6.

7.2 Response of linear/Invariant SystemsFor linear/invariant systems there are four basic single-input, single-response cases,illustrated in Fig. 7.2. They are characterized by the inputs, which are, respectively:

1. a unit impulse at t = 02. a unit step at t = 03. a sinusoid of unit amplitude and frequency f4. white noise

In the first two the system is specified to be quiescent for t < 0 and to be subjected toa control or disturbance input at t = O.In the last two cases the input is presumed tohave been present for a very long time. In these two the system is assumed to be sta-ble, so that any initial transients have died out. Thus in case 3 the response is also asteady sinusoid and in case 4 it is a statistically steady state. We discuss the first threeof these cases in the following, but the fourth, involving the theory of randomprocesses, is outside the scope of this text. The interested reader will find a full ac-count of that topic in Etkin, 1972.

Page 216: Stability Control EtkinReid

208 Chapter 7. Response to Actuation of the Controls-Open Loop

c

o t t----.,:!--~r--~_+,(1)

c

o(2)

c

"A~'-EJ~~\;""""V~ ," (.) -~ V ~(4)

Figure 7.2 The four basic response problems. (1) Impulse response. (2) Step response. (3)Frequency response. (4) Response to white noise.

TRANSFER FUNCTIONS

A central and indispensable concept for response analysis is the transfer function thatrelates a particular input to a particular response. The transfer function, almost uni-versally denoted G(s), is the ratio of the Laplace transform of the response to that ofthe input for the special case when the system is quiescent for t < O. A system with nstate variables Xi and m controls cj would therefore have a matrix of nm transfer func-tions Gij(s).

The Laplace transform of (7.1,4) is

sf = Ai + Be (7.2,1)henceandwhere

(sI - A)i = Be

x =GcG(s) = (sI - A)-IB

(a) (7.22)(b) ,

is the matrix of transfer functions. The response of the ith state variable is then givenby

x;(s) =I G;/s)cls) (7.2,3)j

For a single-input single-response system with transfer function G(s) we have simply

xes) = G(s)c(s) (7.2,4)

Page 217: Stability Control EtkinReid

7.2 Response of Linear/Invariant Systems 209

Figure 7.3 Systems in series.

SYSTEMS IN SERIES

When two systems are in series, so that the response of the first is the input to thesecond, as in Fig. 7.3, the overall transfer function is seen to be the product of thetwo. That is,

andXI(s) = GI(s)C(s)

X2(S) = G2(S)XI(S) = G2(S)GI(S)C(S)

Thus the overall transfer function is

G(s) = xis)/c(s) = GI(s)Gis) (7.2,5)

Similarly for n systems in series the overall transfer function is

(7.2,6)

IDGH·ORDER SYSTEMS

High-order linear/invariant systems, such as those that occur in aerospace practise,can always be represented by a chain of subsystems like (7.2,6). This is important,because the elemental building blocks that make up the chain are each of a simplekind-either first-order or second-order. To prove this we note from the definition ofan inverse matrix (Appendix A.l) that

-I adj (sI - A)(sI - A) = det (sI - A) (7.2,7)

We saw in Sec. 6.1 that det (A - sI) is the characteristic polynomial of the system.We also have, from the definition of the adjoint matrix as the transpose of the matrixof cofactors, that each element of the numerator of the right side of (7.2,7) is also apolynomial in s. (See Exercise 7.1.) Thus it follows from (7.2,2b), on noting that B isa matrix of constants, that each element of G is a ratio of two polynomials, which canbe written as

(7.2,8)

in which f(s) is the characteristic polynomial. It is seen that all the transfer functionsof the system have the same denominator and differ from one another only in the dif-ferent numerators. Since f(s) has the roots Al ... Am the denominator can be factoredto give

(7.2,9)

Page 218: Stability Control EtkinReid

210 Chapter 7. Response to Actuation of the Controls-Open Loop

- .2 + Cln+m' + "n+m-2- -2-

m first-ordercomponents

1/2 (n - m) second-order components

Figure 7.4 High-order systems as a "chain."

Now some of the eigenvalues Ar are real, but others occur in complex pairs, so to ob-tain a product of factors containing only real numbers we rewrite the denominatorthus

m 112(n+m)

f(s) =TI (s - Ar) TI (S2 + a.s + br)r=l r=m+l

(7.2,10)

Here Ar are the m real roots of f(s) and the quadratic factors with real coefficients a;and b; produce the (n - m) complex roots. It is then clearly evident that the transferfunction (7.2,9) is also the overall transfer function of the fictitious system made upof the series of elements shown in Fig. 7.4. The leading component Nij(s) is of courseparticular to the system, but all the remaining ones are of one or other of two simplekinds. These two, first-order components and second-order components, may there-fore be regarded as the basic building blocks of linear/invariant systems. It is for thisreason that it is important to understand their characteristics well-the properties ofall higher-order systems can be inferred directly from those of these two basic ele-ments.

7.3 Impulse ResponseThe system is specified to be initially quiescent and at time zero is subjected to a sin-gle impulsive input

cit) = 8(t)

The Laplace transform of the ith component of the output is then

x;(s) = Gij(s)5(s)

which, from Table (A.1), item 1, becomes

x;(s) = Gi}{s)

This response to the unit impulse is called the impulse response or impulsive admit-tance and is denoted hit). It follows that

his) = Gij(s) (a)

that is, G(s) is the Laplace transform of h(t)

(7.3,1)

Gis) = (' hij(t)e-st dt

From the inversion theorem, (A.2,1l) hit) is then given by

(b) (7.3,2)

Page 219: Stability Control EtkinReid

7.3 ImpulseResponse 211

h ..(t) = _1_ f G..(s)est dsIJ 2m Jc IJ

(7.3,3)

Now if the system is stable, all the eigenvalues, which are the poles of Gij(s) lie in theleft half of the s plane, and this is the usual case of interest. The line integral of(7.3,3) can then be taken on the imaginary axis, s = uo, so that (7.3,3) leads to

(7.3,4)

that is, it is the inverse Fourier transform of Gij(iw). The significance of Gij(iw) willbe seen later.

For a first-order component of Fig. 7.4 with eigenvalue A the differential equa-tion is

x-Ax=c (7.3,5)

for which we easily get

1G(s) = n(s) = --s-A (7.3,6)

The inverse is found directly from item 8 of Table A.1 as

h(t) = e"

For convenience in interpretation, A is frequently written as A = -liT, where T istermed the time constant of the system. Then

h(t) = er'" (7.3,7)

A graph of h(t) is presented in Fig. 7.5a, and shows clearly the significance of thetime constant T.

For a second-order component of Fig. 7.4 the differential equation is

y + 2(w,;j + w~ = c

where x = [y yf is the state vector. It easily follows that

(7.3,8)

_ 1G(s) = h(s) = 2 Y 2

S + 2!owns + Wn

Let the eigenvalues be A = n ± iw, where

n = -(wn

W = wn(l - e)1I2

(7.3,9)

then n(s) becomes

1h(s) = --------

(s - n - iw)(s - n + iw)

1=------,----

(s - nf + w2

(7.3,10)

Page 220: Stability Control EtkinReid

212 Chapter 7. Response to Actuation of the Controls-Open Loop

h(t)

1e

1.0

o,L-------=;-----=---=~::.......-___+

~(t)

Th---r------------/1

TIe / I1_-J_1

/ I/ I

/ Ih I

I

(6)

Figure 7.5 Admittances of a first-order system.

and the inverse is found from item 13, Table A,1 to be

1h(t) = - ent sin wt

w(7.3,11)

For a stable system n is negative and (7.3,11) describes a damped sinusoid of fre-quency to. This is plotted for various, in Fig. 7.6. Note that the coordinates are sochosen as to lead to a one-parameter family of curves. Actually the above result onlyapplies for '::5 1. The corresponding expression for ,~ 1 is easily found by thesame method and is

1h(t) = - e" sinh w't

w'(7.3,12)

where

Graphs of (7.3,12) are also included in Fig. 7.6, although in this case the second-or-der representation could be replaced by two first-order elements in series.

Page 221: Stability Control EtkinReid

7.4 Step-Function Response 213

1.0

I r\V \~ ~

I-r.0.1 r-,~0.5r

~~.0.7 I \1.0

2.0- ~ ~ ./i0002.0 1/ 1\1--1.0

\'\: " -...c;"

\ r-, 0.7

~ \ /0.5

\ / 1'---V\0.1 /.•.•...•

M ~ U ~tII.'/2r

Figure 7.6 Impulsive admittance of second-order systems.

2.0

0.8

0.6

0.4

02

-0.2

-0.4

-0.6

-0.8

I

7.4 Step-Function ResponseThis is like the impulse response treated above except that the input is the unit stepfunction I(t), with transform l/s (Table A.I). The response in this case is called thestep response or indicial admittance, and is denoted dy(t). It follows then that

- - Gy(s)dij(s) = Gij(s)l(s) = --

s(a)

or_71 hij(s)8l.ij(s)= --

S

(7.4,1)

(b)

Since the initial values (at t = 0-) of hi/!) and dy(t) are both zero, the theorem(A.2,4) shows that

dy(t) = J: hij(T) dr (a)

ddy(t)(7.4,2)

or h,P) = (b)• dt

Thus diP) can be found either by direct inversion of (7.4,lb) or by integration ofhy(t). By either method the results for first- and second-order systems are readily ob-tained, and are as follows (for a single input/response pair the subscript is dropped):

Page 222: Stability Control EtkinReid

214 Chapter 7. Response to Actuation of the Controls-Open Loop

2.0

1.8

1/'\

/ \/ \ V""'1\I

/ /" ~ / \-rOIl;; ~ ,," V-:)V --- 'F-- \TIl' "f:);~ --r/!~~~.~1\ I<,

'!'- v'l.S)

-f V \.,VJV

II2.0

,1

1.6

1A

1.2~~ 1.0i

0.8

0.6

0.4

0.2

00 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8wrat/2r

Figure 7.7 Indicial admittance of second-order systems.

First-order system:

d(t) = T(l - e-t/l) (7.4,3)

Second-order system:

d(t) = ~ [1 - ent(cos wt - !!.- sin wt)], t « 1 (7.4,4)Wn W

For' > 1, see Appendix A.2.Graphs of the indicial responses are given in Figs. 7.5b and 7.7.The asymptotic value of d(t) as t ~ 00 is called the static gain K. Applying the

final value theorem (A.2,12) (7.4,1) yields

lim d(t) = lim s.sa(s)= lim G(s)t~lXl ~O s-+O

Thus K = lim G(s) (7.4,5).•...•0

7.5 Frequency ResponseWhen a stable linear/invariant system has a sinusoidal input, then after some time thetransients associated with the starting conditions die out, and there remains only asteady-state sinusoidal response at the same frequency as that of the input. Its ampli-tude and phase are generally different from those of the input, however, and the ex-pression of these differences is embodied in the frequency-response function.

Page 223: Stability Control EtkinReid

7.5 Frequency Response 215

Consider a single input/response pair, and let the input be the sinusoid al cos wt.We find it convenient to replace this by the complex expression c = AIei"'t, of whichal cos wt is the real part. Al is known as the complex amplitude of the wave. The re-sponse sinusoid can be represented by a similar expression, x = A2e

iCdt, the real part

of which is the physical response. As usual, x and c are interpreted as rotating vectorswhose projections on the real axis give the relevant physical variables (see Fig. 7.8a).

From Table A.I, item 8, the transform of c is

Alc= s - uo

1m

x

c Re

(8)

1m, Col

Zero

f

----....,O.•.•...------R.(6)

Figure 7.8 (a) Complex input and response. (b) Effect of singularity close to axis.

Page 224: Stability Control EtkinReid

216 Chapter 7. Response to Actuation of the Controls-Open Loop

Therefore

G(s)i =AI •

S -100

The function G(s) is given by (7.2,9) so that

N(s)i=AI---- (s - ioo)f(s)

The roots of the denominator of the r.h.s. are

(7.5,1)

so that the application of the expansion theorem (A.2,10) yields the complex output

n+1 [ (s - Ar)N(s) ]x(t) = Al L. eA,I

r=1 (s - zoo)f(s) s=Ar

= A [N(ioo) eillJt + c eA1t + c eA2t + ... + c eM]I f(ioo) I 2 n

Since we have stipulated that the system is stable, all the roots Al ••. An of the charac-teristic equation have negative real parts. Therefore eArl ~ 0 as t ~ 00 for r = 1 ... n,and the steady-state periodic solution is

(7.5,2)

N(ioo) .x(t) = A -- e'llJt

] f(ioo) , t~oo

or

x(t) = AIG(ioo)eillJt

=A2eillJt

Thus

A2 = AIG(ioo)

is the complex amplitude of the output, or

(7.5,3)

A2G(ioo) =-

A]the frequency response function, is the ratio of the complex amplitudes. In general,G(ioo) is a complex number, varying with the circular frequency oo.Let it be given inpolar form by

(7.5,4) ,.J

G(ioo) = KMeirp

where K is the static gain (7.4,5). Then

A2 .- =KMe'rp

Al

From (7.5,7) we see that the amplitude ratio of the steady-state output to the input isiAiAII = KM: that is, that the output amplitude is a2 = KMal, and that the phase re-

(7.5,6)

(7.5,7)

Page 225: Stability Control EtkinReid

7.5 Frequency Response 217

1m

O~-- ...•...-----«:::~----~f.tJT=O

Re

f.tJT= 1Figure 7.9 Vector plot of Mi'P for first-order systems.

lation is as shown on Fig. 7.8a. The output leads the input by the angle 'P.The quan-tity M, which is the modulus of G(iw) divided by K, we call the magnification factor,or dynamic gain, and the product KM we call the total gain. It is important to notethat M and 'P are frequency-dependent.

Graphical representations of the frequency response commonly take the form ofeither vector plots of Meitp (Nyquist diagram) or plots of M and 'P as functions of fre-quency (Bode diagram). Examples of these are shown in Figs. 7.9 to 7.13.

EFFECT OF POLES AND ZEROS ON FREQUENCY RESPONSEWe have seen (7.2,9) that the transfer function of a linear/invariant system is a ratioof two polynomials in s, the denominator being the characteristic polynomial. Theroots of the characteristic equation are the poles of the transfer function, and the rootsof the numerator polynomial are its zeros. Whenever a pair of complex poles or zeroslies close to the imaginary axis, a characteristic peak or valley occurs in the ampli-tude of the frequency-response curve together with a rapid change of phase angle atthe corresponding value of w. Several examples of this phenomenon are to be seen inthe frequency response curves in Figs. 7.14 to 7.18. The reason for this behavior isreadily appreciated by putting (7.2,9) in the following form:

G(s) = (s - ZI) • (s - zz) ... (s - zm)(s - AI) . (s - Az) .•• (s - An)

where the Ai are the characteristic roots (poles) and the Zi are the zeros of G(s). Let

(s - Zk) = P#iotk

(s - Ak) = r#ifJk

where p, r, ex,{3are the distances and angles shown in Fig. 7.8b for a point s = uo onthe imaginary axis. Then

Page 226: Stability Control EtkinReid

218 Chapter 7. Response to Actuation of the Controls-Open Loop

1.01\

\

\M ~.....---•... -I{J

\ VX

/ -,""-

/..••..r-,--I -F.::: '--

V

100

go8070

0.8

0.6

M

0.4

0.2 0.4 0.6 0.8IT(0)

1.0 1.2 1.4

.,Q 0'g

~-10

JI"" 0 db/decad~ -~ -26 db/Jecade.•..... ./<,<, r-; r-,

-r--.........•..

...••.•...<; <,-1 2

CAl']'(6)

Figure 7.10 Frequency-response curves-first-order system.

4 10 20 40 100

2~-20

~-30

Os. 0

i-30s1lI-60IIIf_go

0.01 0.02 0.04 0.1 0.2 0.4

When the singularity is close to the axis, with imaginary coordinate w' as illustratedfor point S on Fig. 7.8b, we see that as w increases through to', a sharp minimum oc-curs in p or r, as the case may be, and the angle a or {3increases rapidly through ap-proximately 180°. Thus we have the following cases:

1. For a pole, in the left half-plane, there results a peak in lal and a reduction in({J of about 180°.

2. For a zero in the left half-plane, there is a valley in lal and an increase in ({J ofabout 180°.

3. For a zero in the right half-plane, there is a valley in lal and a decrease in ({J ofabout 180°.

Page 227: Stability Control EtkinReid

7.5 Frequency Response 219

1m

1.0 Re

0.5 = flJ/flJn

Figure 7.11 Vector plot of Meicp for second-order system. Damping ratio ( = 0.4.

FREQUENCY RESPONSE OF FIRST-ORDER SYSTEM

The first-order transfer function, written in terms of the time constant Tis

1G(s) = s + liT

whence

K= limG(s) = T$-+0

5.0

M

fIIIrl\

!t- r-o~0.10

h r/ 1\ ~l,.-0.15~0.20v: j...-. \~

[....-0.30..-0.50

~ ~~VO.70

~or-,r--r-~ v 2.00l"-t--

4.0

3.0

2.0

1.0

oo ~ ~ U U W U U ~flJ/flJn

Figure 7.12 Frequency-response curves-second-order system.

(7.5,8)

Page 228: Stability Control EtkinReid

20

o

If ~r.0.05

It~ ~-

~~~\ 0.25

Q]I\)

"""'" 0.5-.,-....;r-\:::~ ~r.t:fi~

~~IiIl..

f'~ ~ r-,

r>-140db~decaie-

'" '"02 0.3 0.4 0.5 0.6 0.8 1.0w/w"(a)

Figure 7.13 Frequency-response curves-second-order system.

2 3 4 5 6 8 10

10

.a"CI

::i0-10'i

-20

-30

-400.1

Page 229: Stability Control EtkinReid

oL.iL.....:!~~-:~~-:$,..----2l..-~8~~~~~~!----,g!:--...",J2a

I I I I '" I'.Ii> '&llkJe aseqd

,-.."l::l~e"';::ql ~e a-a- '-'

llCl l"'l0 •...•

t-:~

"! :::t0 .ao

~Ind

~0

a

Nd

221

Page 230: Stability Control EtkinReid

222 Chapter 7. Response to Actuation of the Controls-Open Loop

PhugOid: Short- II period:I II ,I II II I

102 '--_----'_ ...•.........•.---L---L_, 1-L.L.L.. __ ....L..---l_~oJ......J.....L..L.a.1 L- __ I...-...L---l.--l-l...l..l.J.J10-2

(J) (rad/s)

(a)

oI- ,~I-Phugoid

approx.

Exact-,lioI-

\ -~f- \~ »->

<,I- \ -,"---- Phugoid: Short- I

oerlod I

I II I

- I I

I I I I dl I I I I I I I I 1:1 I I I I I I II

-60

-120

i -180

-240

-300

-36010-2

(J) (rad/s)

(b)

Figure 7.14 Frequency-response functions, elevator angle input. Jet transport cruising at highaltitudes. (a) Speed amplitude. (b) Speed phase.

The frequency response is determined by the vector G(iw)

. TG(iw) = KMe"P = ---

1+ iwT

whence

(7.5,9)

Page 231: Stability Control EtkinReid

7.5 Frequency Response 223

Short-periodapprox,

""

102 b-----------+---------+------'''''------iPhugoid: Short- :

: period:I 11 1

1011-__ L----L---l---l......J..:.L1 ..l-J..L- __ l...----I.----J----J'--L....L.J~I...J..._ __ _'______l_J.._.L-J._'_ •..••.•10-2 10-1 100 10'

coIrad/s)(a)

o~ Phugoid

approx.

0

- Exact",,/\0

j.\'-

0

~ Short-period7---~approx.

0

Short\~ Phugoid:I period: ~0I r <;I 1~ I 1

I I I 1 1:1 II 1 I 1 I I 1 1:1 1 I 1 I I I 11

-6

-12

"'~r;F -18

-24

-30

-36010-2

co(rad/s)(b)

Figure 7.15 Frequency-response functions, elevator angle input. Jet transport cruising at highaltitude. (a) Angle of attack amplitude. (b) Angle of attack phase.

From (7.5,9), M and 'P are found to be1

M = (1 + w2T2)1I2 (7.5,10)

'P = -tan-l wTA vector plot of Mei<p is shown in Fig. 7.9. This kind of diagram is sometimes calledthe transfer-function locus. Plots of M and 'P are given in Figs. 7.lOa and b. The ab-scissa is fT or log wT where f = W/21T, the input frequency. This is the only parame-

Page 232: Stability Control EtkinReid

224 Chapter 7. Response to Actuation of the Controls-Open Loop

ro (rad/s)(a)

o -~ "- Exact --"" Phugoid

approx.~

Short-period I\. ~"- approx. ~ ______ 'Sr\

f-

~Phuqoid] Short-: "----

f-I periodjI 1

I I

'- I II I

I I I I III II I I I I I 11:1 I I I I I I II

-6

-12

••~ -18r;'

-24

-30

-3610-2

ro (rad/s)

(b)

Figure 7.16 Frequency-response functions, elevator angle input. Jet transport cruising at highaltitude. (a) Pitch-rate amplitude. (b) Pitch-rate phase.

ter of the equations, and so the curves are applicable to all first-order systems. Itshould be noted that at Cd = 0, M = I and ({J = O.This is always true because of thedefinitions of K and G(s)-it can be seen from (7.4,5) that G(O) = K.

FREQUENCY RESPONSE OF A SECOND-ORDER SYSTEM

The transfer function of a second-order system is given in (7.3,9). The frequency-re-sponse vector is therefore

Page 233: Stability Control EtkinReid

7.5 Frequency Response 225

OJ (rad/s)

(a)

-6

0

I-

~~ :\~

~ Exact/:I

IPhugoidI

I- Short-period IV approx.approx. h

\ I ,~ II ---- II --- I Short-

PhugOid: -rfo- II I I I I ~I I I I I I I I I I I I I I I I II

-12

1: -18e-

-24

-30

-3610-2

OJ (rad/s)

(b)

Figure 7.17 Frequency-response functions, elevator angle input. Jet transport cruising at highaltitude. (a) Flight-path angle amplitude. (b) Flight-path angle phase.

w~Mei'P = --::-----=------(w; - (2) + 2i(wnw

From the modulus and argument of (7.5,11), we find that

1M=-----=-=----:-----:-_:_=_{[I - (WIWn)2]2 + 4e(w1wn?}112

_ -1 2(wlWn

cp - -tan 1 - (w1wn?

(7.5,11)

(7.5,12)

Page 234: Stability Control EtkinReid

226 Chapter 7. Response to Actuation of the Controls-Open Loop

102 i=""------~o:----r---------~--------___.

Short-periodapprox.

101 E====~~t====l:=~~:;;;;;.-~;"'--P~--------J

10° F-~f--------t----------t-------"::~::----I

w (rad/s)(a)

Phugoid: Short- I

I period:I II II I

1O-1'---:------'----'----'----'---'-'1L..L-.L.L __ ...L-_.L..-..l...-.J.....J-.L.L1:L..L __ ....L_L-...L-..L-L....L..LJ..J

1~ 1~ 1~

o ....I- Exact __ 30"\

\,,

'- \\Short-period \.\ __ Phugoid

I-approx. "" approx.

\ '--, ---

~I

Short~- Phugoid:I period I 1\I I <;- I II I

I I I I ,:, I I I I I I II d

-6

-12

i -18

-24

-30

-3610-2

co (rad/s)(b)

Figure 7.18 Frequency-response functions, elevator angle input. Jet transport cruising at highaltitude. (a) Load factor amplitude. (b) Load factor phase.

A representative vector plot of Mei"" for damping ratio C = 0.4, is shown in Fig. 7.11,

and families of M and cp are shown in Figs. 7.12 and 7.13. Whereas a single pair ofcurves serves to define the frequency response of all first-order systems (Fig. 7.10), ittakes two families of curves, with the damping ratio as parameter, to display the char-acteristics of all second-order systems. The importance of the damping as a parame-ter should be noted. It is especially powerful in controlling the magnitude of the reso-nance peak which occurs near unity frequency ratio. At this frequency the phase lagis by contrast independent of C, as all the curves pass through cp = -900 there. For all

Page 235: Stability Control EtkinReid

7.5 Frequency Response 227

values of (, M ~ 1 and ({)~ 0 as w/wn ~ O. This shows that, whenever a system isdriven by an oscillatory input whose frequency is low compared to the undampednatural frequency, the response will be quasistatic. That is, at each instant, the outputwill be the same as though the instantaneous value of the input were applied stati-cally.

The behavior of the output when (is near 0.7 is interesting. For this value of (, itis seen that ({)is very nearly linear with w/wn up to 1.0. Now the phase lag can be in-terpreted as a time lag, T = «({)/2n)T = ({)/w where T is the period. The output waveform will have its peaks retarded by T sec relative to the input. For the value of (un-der consideration, ({)/(w/wn) =i= 'Tr/2or ({)/w= n/2wn = tTm where T; = 2'Tr/wn,the un-damped natural period. Hence we find that, for ( =i= 0.7, there is a nearly constanttime lag T =i= tTn, independent of the input frequency, for frequencies below reso-nance.

The "chain" concept of higher-order systems is especially helpful in relation tofrequency response. It is evident that the phase changes through the individual ele-ments are simply additive, so that higher-order systems tend to be characterized bygreater phase lags than low-order ones. Also the individual amplitude ratios of the el-ements are multiplied to form the overall ratio. More explicitly, let

G(s) = GI(s) . G2(s) ... Gn(s)

be the overall transfer function of n elements. Then

G(iw) = GI(iw) . G2(iw) ... GnCiw)= (KIM! . K2M2 ••• KPn)ei('Pl+fP2+"''Pn)

= KMeirp

so that (a)

n({)= I ({)r

r=l

(7.5,13)(b)

On logarithmic plots (Bode diagrams) we note thatn

log KM = I log s.s«.r=l

(7.5,14)

Thus the log of the overall gain is obtained as a sum of the logs of the componentgains, and this fact, together with the companion result for phase angle (7.5,13)greatly facilitates graphical methods of analysis and system design.

RELATION BETWEEN IMPULSE RESPONSE AND FREQUENCYRESPONSE

We saw earlier (7.3,4), that h(t) is the inverse Fourier transform of G(iw), which wecan now identify as the frequency response vector. The reciprocal Fourier transformrelation then gives

G(iw) = f' h(t)e-iwr dt-00

(7.5,15)

Page 236: Stability Control EtkinReid

228 Chapter 7. Response to Actuation of the Controls-Open Loop

that is, the frequency response and impulsive admittance are a Fourier transformpair.

7.6 Longitudinal ResponseTo treat this case we need the matrices A and B of (7.1,4). A is given in (4.9,18), butwe have not yet given B explicitly. On the right side of (4.9,18) we have the productB.dc given by

B.dc = (m - 2.;,)

tiMe Mw AZe--+-----Iy Iy (m - Zw)

o

sx,maz,

(7.6,1)

We now have to specify the control vector c and the corresponding aerodynamicforces and moment. For longitudinal control, we assume here that the available con-trols are well enough represented by

(7.6,2)

and that the incremental aerodynamic forces and moment that result from their actua-tion are given by a set of control derivatives XSe and so on, in the form

(7.6,3)

Additional elements can be added to c and to (7.6,3) if the situation requires it.The use of constant derivatives, as in (7.6,3), to describe the force output of the

propulsion system in response to throttle input does not allow for any time lag in thebuildup of engine thrust since it implies that the thrust is instantaneously proportionalto the throttle position. This is not unreasonable for propeller airplanes, but it is not agood model for jets in situations when the short-term response is important, as for ex-ample in a balked landing. To allow for this effect when the system is modeled in theLaplace domain, one can use control transfer functions instead of control derivatives.That is, one can replace, for example, X8p by Gxap(s). If the system model is in thetime domain, the same result can be obtained by adding an additional differentialequation and an additional variable. This latter method is illustrated in the example ofSec. 8.5.

By substituting (7.6,3) into (7.6,1) we derive the matrix B to be

Page 237: Stability Control EtkinReid

7.7 Responses to Elevator and Throttle 229

Xae Xapm m

Zae ZapB= (m - Zw) (m - Zw) (7.6,4)

M MwZae M MwZap~+ ~+t, lim - Zw) t, ly(m - Zw)

0 0

With A and B known, we can compute the desired transfer functions and responses.(In this example the elevator angle is in radians, and English units are used for allother quantities). We calculate responses for the same jet transport as was used previ-ously in Sec. 6.2, with A given by (6.2,1). The nondimensional elevator derivativesare:

CX6e

= -3.818 X 10-6

CZ6e= -0.3648Cm6e = -1.444

from which the dimensional derivatives are calculated as

Xa = CX6 ~pu~S = -3.717e e

Zae = CZ6e~pu~S = -3.551 X 105

Mae = Cm6e~pu~Sc = -3.839 X 107

For the throttle, we arbitrarily choose a value of Xa/m = 0.3 g when 5p = 1, and Zapand Map = O.With these values we get for the matrix B:

[

-0.000187 ~90.66]B = -17.85

-1.158o

(7.6,5)

7.7 Responses to Elevator and ThrottleRESPONSE TO ELEVATOR

When the only input is the elevator angle A5e the system reduces to

[AU] [bll]

(sf - A) ~ ~ :~: AB. (7.7,1)

where bij are the elements of B. Solving for the ratios Guae(s) = AiilAlJe and so onyields the four transfer functions.' Each is of the form (7.2,8). For this case the char-

lIn the subscripts for the transfer function symbols, the symbol d is omitted in the interest of sim-plicity.

Page 238: Stability Control EtkinReid

230 Chapter 7. Response to Actuation of the Controls-Open Loop

acteristic polynomial is the left side of (6.2,2) (with s replacing A), and the numeratorpolynomials are

Nu8e = -0.000188s3- 0.2491s2 + 24.68s + 11.16

Nw8e= -17.85s3- 904.0s2

- 6.208s - 3.445Nq8e = -1.158s3 - 0.3545s2 - 0.003873sN98e = -1.158s2

- 0.3545s - 0.003873

There are two other response quantities of interest, the flight path angle I' and theload factor nz• Since eo = 0, Ae = e, and AI' = AO - Aa, it readily follows that

(7.7,2)

GY8e = G98e - G"'8e (7.7,3)

We define nz to be (see also Sec. 3.1)

(7.7,4)

It is equal to unity in horizontal steady flight, and its incremental value during the re-sponse to elevator input is

Anz = -AZlW = -(ZuAu + Zww + Zqq + Zww + Z8.A5e)/W

After taking the Laplace transform of the preceding equation and dividing by A5e, weget the transfer function for load factor to be

••

(7.7,5)

The total gain and phase of the frequency responses calculated by (7.5,6) from five ofthe above six transfer functions (for Au, w, q, AI', and Anz) are shown in Figs. 7.14 to7.18.

The exact solutions show that the responses in the "trajectory" variables u and I'are dominated entirely by the large peak at the low-frequency Phugoid mode. Be-cause of the light damping in this mode, the resonant gains are very large. The peakIGu8.1 of nearly 3 X 104 means that a speed amplitude of 100 fps would result from anelevator angle amplitude of about 100/(3 X 104) rad, or about 0.2°. Similarly, at reso-nance an amplitude of 10° in I' would be produced by an elevator amplitude of aboutiO. For both of these variables the response diminishes rapidly with increasing fre-quency, becoming negligibly small above the short-period frequency. The phase an-gle for u, Fig. 7.14b, is zero at low frequency, decreases rapidly to near -180° at thephugoid frequency (very much like the lightly damped second-order systems of Fig.7.13) and subsequently at the short-period frequency undergoes a second drop char-acteristic of a heavily damped second-order system. The "chain" concept of high-or-der systems (Sec. 7.2) is well exemplified by this graph.

By contrast, the attitude variables w and q show important effects at both low andhigh frequencies. The complicated behavior of w near the phugoid frequency indi-cates the sort of thing that can happen with high-order systems. It is associated with apolelzero pair of the transfer function being close to one another. Again, above theshort-period frequency, the amplitudes of both w and q falloff rapidly.

The amplitude of the load factor Anz has a very large resonant peak at thephugoid frequency, almost loo/rad. It would not take a very large elevator amplitudeat this frequency to cause structural failure of the wing!

Page 239: Stability Control EtkinReid

7.7 Responses to Elevator and Throttle 231

Step-Function ResponseThe response of the airplane to a sudden movement of the elevator is shown by

the step response. This requires a solution in the time domain as distinct from the pre-ceding solution, which was in the frequency domain. Time domain solutions are com-monly obtained simply by integrating (7.1,4) by a Runge-Kutta, Euler, or other inte-gration scheme, the choice being dependent on the order of the system, accuracyrequired, computer available, and so on. The software used" for the example to followdoes not integrate the equations, but instead uses an alternative method. It inverts thetransfer function using the Heavyside expansion theorem (A,2,10). For the same jetairplane and flight condition as in the preceding example, the control vector for ele-vator input is c = [AcSe O]T and A and B are as before. (Note that only the first col-umn of B is needed.) Time traces of speed, angle of attack, and flight path angle areshown in Figs. 7.19 and 7.20 for two time ranges when the elevator displacement isone degree positive, that is, down.

It is seen from Fig. 7.19, which shows the response during the first 10 sec, thatonly the angle of attack responds quickly to the elevator motion, and that its variationis dominated by the rapid, well-damped short-period mode. By contrast, the trajec-tory variables, speed, and flight path angle, respond much more slowly. Figure 7.20,which displays a 10 min time span, shows that the dynamic response persists for avery long time, and that after the first few seconds it is primarily the phugoid modethat is evident.

The steady state that is approached so slowly has a slightly higher speed and aslightly smaller angle of attack than the original flight condition-both changes thatwould be expected from a down movement of the elevator. The flight path angle isseen to be almost unchanged-it increased by about one-tenth of a degree. The reasonfor an increase instead of the decrease that would be expected in normal cruising flightis that at this flight condition the airplane is flying below its minimum-drag speed.

If the reason for moving the elevator is to establish a new steady-state flight con-dition, then this control action can hardly be viewed as successful. The long lightlydamped oscillation has seriously interfered with it. Clearly, longitudinal control,whether by a human or an automatic pilot, demands a more sophisticated control ac-tivity than simply moving it to its new position. We return to this topic in Chap. 8.

Phugoid ApproximationWe can get an approximation to the transfer functions by using the phugoid ap-

proximation of Sec. 6.3. The differential equation is (6.3,6) with control terms added,that is,

XlSe

Au Au mW W ZlSe=A + ss, (7.7,6)0 q mAiJ AO MISe

0

2ProgramCC

Page 240: Stability Control EtkinReid

232 Chapter 7. Response to Actuation of the Controts-Open Loop

901--------------- --,

60 f-

lJ).E-;:l<l

30 f-

-0 I

0 2 4 6 8 10Time, S

(a) Speed

0

-.01

"0~tl<l

-.02 •

-.030 12 24 36 48 60

Time,s(b) Angle of attack

.1

.08

.06

.04

.02"0~ 0~

-.02

-.04

-.06 ••-.08

-.10 2 4 6 8 10

Time, S

(e) Flight path angle

Figure 7.19 Response to elevator (~Be = 10). Jet transport cruising at high altitude.

Page 241: Stability Control EtkinReid

90,-----------------------------,

7.7 Responses to Elevator and Throttle 233

60

30~u•• = 46.3 fps

120 360 600480240Time,s

(a) Speed

•.0...---------------------.----------,

-.01

-.02

-.030 120 240 360 480 600

Time, S

(bl Angle of attack

.1

.08

.06

.04

.02"01': 0~

-.02

-.04

-.06

-.08

120 240 360 480 600Time. S

(c) Flight path angle

Figure 7.20 Response to elevator (.18. = 1°). Jet transport cruising at high altitude.

Page 242: Stability Control EtkinReid

where A is the matrix of (6.3,6). After taking the Laplace transform of this equationand solving for the ratios of the variables, we find the transfer functions to be

als + £loG = ~--=-

u8. f(s)

b2s2 + bIs + boG - ~----=--~w8. - f(s)

c2s2 + CIS + CoG - ---=--------"----=-68. - f(s)

where f(s) is the characteristic polynomial of (6.3,8),

f(s) = As2 + Bs + C

234 Chapter 7. Response to Actuation of the Controls-Open Loop

and

(a)

(b) (7.7,7)

(c)

(a)

(b)

••(c)

(d)

(e) (7.7,8)

if)

(g)

(h)

Short-Period ApproximationWe can also get useful approximations to the transfer functions for e, q, and a by

using the short-period approximation of Sec. 6.3. Instead of (7.7,1) we get the equa-tion

(sI - A) [ ; ] = B 6.Se

in which A is the matrix of (6.3,13) and B is obtained from (7.6,4) as

B= mMa• My" 2a•-+--Iy Iy m

(7.7,9)

(7.7,10)

Page 243: Stability Control EtkinReid

7.7 Responses to Elevator and Throttle 235

When calculating the transfer functions we take note of the fact that ij = sAO. Aftersolving (7.7,9) for the appropriate ratios we get the results

(a)

bIs + bo Gq8eG08• = --- = --sf(s) s

f(s) = S2 + CIS + Co

(b) (7.7,11)

(c)

where

Z8eal =--muo

M8e Mq Z8ea -u ----0- 0 I I m

y y

M8 M· Z8bI=-e +~_e

Iy t, m

Z8e u; z, M8ebo=-----m Iy m Iy

(Zw u, u; )

CI = - - + - + - uom Iy t,Mw u, z,

Co= -uo- + --t, t, m

The frequency responses calculated with the above approximate transfer func-tions are shown on Figs. 7.15 to 7.18. It is seen that the phugoid approximation is ex-act at very low frequencies and the short-period approximation is exact in the high-frequency limit. For frequencies between those of the phugoid and short-periodmodes, one approximation or the other can give reasonable results.

(a)

(b)

(C)

(d) (7.7,12)

(e)

if)

RESPONSE TO THE THROTTLE

For the same jet airplane and flight condition as in previous numerical examples, wecalculate the response to a step input in the throttle of ASp = t which corresponds toa thrust increment of 0.05W.The matrix B is given by (7.6,5) and Ac is

Ac = [0 1I6f (7.7,13)

The numerical results are shown in Fig. 7.21. Because the model has not includedany engine dynamics, the results are not valid for the first few seconds. However, thisregion is not of much interest in this case. The motion is clearly seen to be dominatedby the lightly damped phugoid. The speed begins to increase immediately, before theother variables have time to change. It then undergoes a slow damped oscillation, ul-timately returning to its initial value. The angle of attack varies only slightly, and 'Ymakes an oscillatory approach to its final positive value 'Yss' The new steady state is aclimb with Au = Aa = O.When the thrust line does not pass through the CG, the re-sponse is different in several details. Principally, the moment of the thrust causes a

Page 244: Stability Control EtkinReid

236 Clulpter 7. Response to Actuation of the Controls-Open Loop

40

30

20

10U>.e-~

0

-10

-20

-30

-400 600

.05r---------------------------......,

.025 ,....

•"~ o~_ -t$-e

••-.25 f-

I I I I I-.050 100 200 300 400 500 600Time,s

(b) Angle of attack

.2

.15

.05•••

300Time, S

(e) Flight path angle

Figure 7.21 Response to throttle. Jet transport cruising at high altitude. Thrust line passingthroughCG.

100 200 400 500 600

Page 245: Stability Control EtkinReid

7.8 Lateral Steady States 237

rapid change in a, followed by an oscillatory decay to a new Aass =1=0, and the speedconverges to a new value Auss =1=O.

7.8 Lateral Steady StatesThe basic flight condition is steady symmetric flight, in which all the lateral variables{3,p, r, cf>are identically zero. Unlike the elevator and the throttle, the lateral controls,the aileron and rudder, are not used individually to produce changes in the steadystate. This is because the steady state values of {3,p, r, cf> that result from a constant5a or 5r are not generally of interest as a useful flight condition. There are two lateralsteady states that are of interest, however, each of which requires the joint applicationof aileron and rudder. These are the steady sideslip, in which the flight path is recti-linear, and the steady turn, in which the angular velocity vector is vertical. We lookinto these below before proceeding to the study of dynamic response to the lateralcontrols.

THE STEADY SIDESLIP

The steady sideslip is a condition of nonsymmetric rectilinear translation. It is some-times used, particularly with light airplanes, to correct for cross-wind on landing ap-proaches. Glider pilots also use this maneuver to steepen the glide path, since the UDratio decreases due to increased drag at large {3.In this flight condition all the timederivatives in the equations of motion (except xE) and the three rotation rates p, q, rare zero. It is simplest in this case to go back to (4.9,2)-(4.9,6), from which we derivethe following:

AY + mgcf> cos 80 = 0M.=O

AN= 0(7.8,1)

We use (4.9,17) for aerodynamic forces, and for the control forces use the followingas a reasonable representation:

(7.8,2)

We now add the assumption that 80 is a small angle and get the resulting equation

[

Ya 0L~ LaaNaT Naa

(7.8,3)

In this form, v is treated as an arbitrary input, and (5,., 5a, cf» as outputs. (See Exercise7.6.) Clearly, there is an infinity of possible sideslips, since v can be chosen arbitrar-ily. Note that the other three variables are all proportional to v. We illustrate thesteady sideslip with a small general aviation airplane" of 30-ft (9.14 m) span and a

3Based on the Piper Cherokee. The control derivatives were taken from McCormick (1979). We es-timated the stability derivatives. The numerical values used may not truly represent this airplane.

Page 246: Stability Control EtkinReid

238 Chapter 7. Response to Actuation of the Controls-Open Loop

Table 7.2Nondimensional Derivatives-General Aviation Airplane (expressed in rad-1 and (rad/s)-l)

Cy Cj Cn

f3 -0.14 -0.0689 - 0.0917CL 0.01326 + O.017CZ

P -0.039 -0.441 -0.00109 - 0.0966CL

f 0.165 -0.0144 + 0.271CL -0.048 - 0.0238cI

s: 0 -0.0531 0.005

8r 0.117 0.0105 -0.0509

gross weight of 2400 lb (10,675 N). The altitude is sea level and CL = 1.0, corre-sponding to a speed of 112.3 fps (34.23 m/s), and the wing area is 160 ft2 (14.9 nr'),The nondimensional derivatives are given in Table 7.2, from which the numericalsystem equation is found from (7.8,3) to be (see Exercise 7.5)

[

280.7755.7

-3663.5

o-3821.9

359

2400] [5r] [2.991]o 5a = 102.93 v

o ep -19.394(7.8,4)

••

It is convenient to express the sideslip as an angle instead of a velocity. To do so werecall that {3 == v/uo, with Uo given above as 112.3 fps. The solution of (7.8,4) isfound to be

5/{3 = .3035a/{3 = -2.96ep/{3= .104

We see that a positive sideslip (to the right) of say 100 would entail left rudder of 3°and right aileron of 29.6°. Clearly the main control action is the aileron displacement,without which the airplane would, as a result of the sideslip to the right, roll to theleft. The bank angle is seen to be only I° to the right so the sideslip is almost flat.

THE STEADY TURNWe define a "truly banked" tum to be one in which (1) the vehicle angular velocityvector to is constant and vertical (see Fig. 7.22) and (2) the resultant of gravity andcentrifugal force at the mass center lies in the plane of symmetry (see Fig. 7.23). Thiscorresponds to flying the tum on the tum-and-bank indicator," It is quite common forturns to be made at bank angles that are too large for linearization of sin ep and cos epto be acceptable, although all the state variables other than ep and V are small. Thuswe tum to the basic nonlinear equations in Sec. 4.7 for this analysis. The large bankangle has the consequence that coupling of the lateral and longitudinal equations oc-curs, since more lift is needed to balance gravity than in level flight. Thus not onlythe aileron and rudder but the elevator as well must be used for turning at large ep. •.

"Neglecting the fact that the pilot and indicator are not right at the CG.

Page 247: Stability Control EtkinReid

7.8 Lateral Steady StIltes 239

Helical path

Figure 7.22 Steady climbing tum.

(b+-------~ .c

xz plane/

mg

y

Figure 7.23 Gravity and acceleration in tum.

Page 248: Stability Control EtkinReid

240 Chapter 7. Response to Actuation of the Controls-Open Loop

The body-axis angular rates are given by

(7.8,5)

which for small elevation angle (j yields

[:] = [~: </>] cor cos </>

We now apply the second condition for a truly-banked turn, that is, that the ballbe centered in the turn-and-bank indicator. This means that the vector mg - mac,where a, is the acceleration vector of the eG, shall have no y component. But mac isthe resultant external force f, so that from (4.5,6)

(7.8,6)

mg-mac=mg-f= -A

where A is the resultant aerodynamic force vector. Thus we conclude that the aerody-namic force must lie in the xz plane, and hence that Y = O. We consider the casewhen there is no wind, so that

(uE, vE, ~) = (u, v, w)

and choose the body axes so that ax = w = O. We now use (4.7,1) with all the vari-ables constant and only u and </>not small to get:

Y = -mg sin </>+ mru = 0 (a) (7.87),/., (b) ,Z = -mg cos 'I' - mqu

When v is small, a reasonable assumption for a truly banked tum, we also have thatu == V, the flight speed. It follows from (7.8,7a) that

rVsin </>= -

g

and with the value of r obtained from (7.8,6)

coVtan </>= -

g(7.8,8)

The load factor nz is obtained from (7.8,7b):

Z qVn = - - = cos</>+ -

z mg g

With q from (7.8,6) this becomes

Vco sin </>nz= cos </>+ ------:...

g

By using (7.8,8) to eliminate Vco we get

nz = sec </> (7.8,9)

\

Page 249: Stability Control EtkinReid

7.8 LateralSteadyStates 241

We note from (5.1,1) that Z = -L in this case, so that n = LIW = nz• The incremen-tal lift coefficient, as compared with straight flight at the same speed and height, is

L-mgdCL = ip~S = (n - I)Cw (7.8,10)

We can now write down the equations governing the control angles. From (4.7,2), toftrst order, L = M = N = 0, so we have the five aerodynamic conditions

CI = Cm = C; = C; = 0dCL = (n - I)Cwand

On expanding these with the usual aerodynamic derivatives, we get

Clf3{3 + Cli) + Czl' + Cl8r5r + Cl8as, = 0

Cm"da + Cmqq + Cm8ed5e = 0Cnf3{3 + CnpP + c.r + Cn8r5r + Cn8a5a = 0

Cyf3{3 + Cy~ + Cy/' + CY8rs, = 0CL"da + CLqq + CL6ed5e = (n - I)Cw

In these relations P, q, r are known from (7.8,6), that is,

(7.8,11)

bP -0-

2V

Cq sin cf> 2V w (7.8,12)

bf cos cf> 2V

The five equations (7.8,11) for the five unknowns [{3,5r, 5aland [da, d5eluncoupleinto two independent sets:

(7.8,13)

and

[Cm" <;-,[da] [Cmq] we. [ 0 ]=- -smcf>+CLa CL8e_ d5e CLq 2V (n - l)Cw

When (7.8,9) is used to eliminate cf> from (7.8,14), and after some routine algebra, thesolution for d5e is found to be (see Exercise 7.6)

(7.8,14)

(7.8,15)

Page 250: Stability Control EtkinReid

242 Chapter 7. Response to Actuation of the Controls-Open Loop

Except for far forward CG positions and low speeds, the angles given by (7.8,15) aremoderate. The similarity of this expression to that for elevator angle per g in a pull-up (3.1,6a) should be noted. They are in fact the same in the limit n ~ 00. The eleva-tor angle per g in a turn is therefore not very different from that in a vertical pull-up.

Finally, the lateral control angles are obtained from the solution of (7.8,13).

NUMERICAL EXAMPLE

The rudder and aileron angles in a steady truly-banked turn are calculated by way ofexample for the same general aviation airplane as was used above for the sideslip.The altitude is sea level, the speed is 125 fps (38.1 m/s) and the stability and controlderivatives are as in Table 7.2. The solution of (7.8,13) for climb angles between-10° and +10° shows that the sideslip angle f3 remains less than 1.5° and the rudderand aileron angles are as shown in Figs. 7.24 and 7.25. The value of CL varies overthe range of bank angles used from 0.8 to 1.6, so several of the stability derivativesare significantly affected. It is seen that the aileron angle is always positive for a rightturn-that is, the right aileron is down (stick to the left), and that the rudder is usuallynegative (right rudder) although its sign may reverse in a steep climb. The strong ef-fect of the climb angle derives from the fact that the roll rate p is proportional to (Jand changes sign with it. Thus the terms C1jJ and Cnl) in the moment equationschange sign between climbing and descending and affect the control angles requiredto produce zero moment.

••

10

8 -

6 -

4 -0.0- 2 -oje;,<: 0'"(;;"0"0 -2 -:Ja::

-4 -

-6-

-8-

-100

.'100climb .•'..,..............---- ----------------- ----- •..~~~~

10° descent '" ..•.......,

" "'.

Horizontal-

I I I I I

10 20 30 40Bank angle. ~o

50 60

Figure 7.24 Rudder angle in tum.

• General aviation airplane• Speed 125 fps• Sea level

\

Page 251: Stability Control EtkinReid

7.9 Lateral Frequency Response 243

20.----------------..,....-------...,

6

, 100descent,,,,,I,,,

I

: HorizontalIIIIIIIIII

II

II,

",,'~~ .

l

18

16

14

4

10°climb2

Figure 7.25 Aileron angle in tum.

• General aviation airplane• Speed 125fps• Sea level

Finally, it may be remarked that the control angles obtained would have beensubstantially different had it been stipulated that {3,not Cy, should be zero in the turn.It would not then be possible, however, to satisfy the requirement that the ball be cen-tered in the turn-and-back indicator.

7.9 Lateral Frequency ResponseThe procedure for calculating the response of the airplane to sinusoidal movement ofthe rudder or aileron is similar to that used for longitudinal response in Sec. 7.6. The

Table 7.3Control Derivatives-B747 Jet Transport (expressed in rad ")

Cy C1 Cn

s, 0 -1.368 X 10-2 -1.973 X 10-4

8, 0.1146 6.976 X 10-3 -0.1257

Cz

-3.818 X 10-6 -0.3648 -1.444

Page 252: Stability Control EtkinReid

244 Chapter 7. Response to Actuation of the Controls-Open Loop

aerodynamics associated with the two lateral controls are given by a set of control de-rivatives:

"[LlYe] [Y8 Y8 ]si; = L8: L8: [ ::]

ANe N8a N8r

The Laplace transform of the system equation (4.9,19) is then

(7.9,1)

\

(7.9,2)

where B is...

Y8a Y8r

m m

L L~ + 1'zxN ~ +l' NB= l' e, l' zx s;x x

N Nl'L + ~ l'L + ~zx s, l' zx s; l'

z z

0 0

(7.9,3)

NUMERICAL EXAMPLE

For our numerical example we use the same jet transport and flight condition as inSec. 7.6. A is given by (6.7,1) and the control derivatives are given in Table 7.3, fromwhich, with the definitions of Table 7.1, the elements of B are calculated to be

[

0 5.642 ]B - -0.1431 0.1144

0.0~741 -0.~859(7.9,4)

The eight transfer functions are then as in (7.2,8), where f(s) is the characteristicpolynomial of (6.7,2) (with s instead of A)and with the numerators as follows:

Nu8a = 2.896s2 + 6.542s + 0.6220 (a)Nu8r = 5.642s3 + 379.4s2 + 167.9s - 5.934 (b)Np8a = 0.1431s3 + 0.02730s2 + 0.1102s (c)NP8r = 0.1144s3 - 0.1997s2 - 1.368s (d)Nr8a = -0.003741s3 - 0.002708s2 - 0.OOO1394s+ 0.004539 (e) (7.9,5)Nr8r = -0.4859s3 - 0.2327s2

- 0.009018s - 0.05647 if)N</>8a = .1431s2 + 0.02730s + 0.1102 (g)

N</>8r = 0.1144s2- 0.1997s - 1.368 (h)

Page 253: Stability Control EtkinReid

7.9 Lateral Frequency Response 245

From the transfer functions Gij(s) = Nij(s)/f(s), the frequency response functionsGiiw) were calculated for both aileron and rudder inputs. The results for D, l/J and rare shown on Figs. 7.26 and 7.27. The most significant feature in all of these re-sponses is the peak in the amplitude at the Dutch Roll frequency, and the associatedsharp drop in phase angle.

At zero frequency we see from (7.9,5c and d) that the roll rate amplitude is zerofor both inputs. All the other variables have finite values at w = O.Even for moderate

Spiral/rollapprox.~

1021---------~L---------____f-----.:~-----__1I,

Dutch:roll I,,

III,

llJ (rad/s)(a)

180

90I'-....<...

/ ~Piral/rollapprox.

------- \---", I.-- Exact

DutCh! \roll approx. \

\III

II

IDutch: r\.roll'

I I I I I I II I I I I I I III

i 0

-90

-18010-2

llJ (rad!s)

(b)

Figure 7.26 Frequency-response functions, rudder angle input. Jet transport cruising at highaltitude. (a) Sideslip amplitude. (b) Sideslip phase. (c) Roll amplitude. (d) Roll phase. (e) Yaw-rateamplitude. if) Yaw-rate phase.

Page 254: Stability Control EtkinReid

246 Chapter 7. Response to Actuation of the Controls-Open Loop

1021=""---........::---------,r------- __ ---. ---,

"''''''''''''''''.•.

"' .•.Spiral/roll ~ .••approx.

10' F=----------1------:O~;:__---f-Ir_-----------1

10° F----------f-----------+~~-------~.•.,.•.,.•..•.,.•..•.,,

10-1 L....:---L----I.-L-L..J......LL.l..J __ ----J'----l.----J----J--L...LJ-l.lJ. __ ___l.l.--l..~___l....L.Ll..l..J10-2

co(rad/s)(c)

180r---------...,-------------,-----------,

90t-------=-.....,J;:::-----------+------------I

,

Spiral/roll~~~~ ~~ . .-/ approx .

•..••.•..•..••.$- ..._-------------! Ot------------l----------++---------~

-9a-----------+-------------I't------------i

co(rad/s)

(d)

Figure 7.26 (Continued)

control angles, however, the steady-state values of {3= vluo and cP are very large (seeExercise 7.10). Hence the linearity assumption severely constrains these zero fre-quency solutions. If, however, we postulate that the control angles are so small thatthe linearity conditions are met, then there is a steady state with constant values of cP,{3,and r. This can only be a horizontal turn in which the angular velocity vector is

o, = [0 qss rssfwhere qss = n sin cP

rss = ncos cP

,

Page 255: Stability Control EtkinReid

7.10 Approximate Lateral Transfer Functions 247

101 r-----------.,-----------..,.------------,

10° I----------....f'o-~-------_H_-+;_-----------j

10-1 1:::----------+-7"''''---------....:..;:-t--~~~--'''''''=_-"''''~ Spiral/roll

, ••. /approx.I "'••.

Dutch I •••

roll: "''''••.: " ...I ••••••I ••••••I ••••••

co(rad/s)(e)

90,---------F'""==:-------,------------,••••••••• Spiral/roll•...

••••••••••••lapprox .•.. •.•...•.•.•.•. _-----ol----------I-------\----+----~= •.......----1

"'~~-9d-----------+-----------'~---------____jDutch

roll approx,

-181d-----------+------------tr-----------

co(rad/s)

(flFigure 7.26 (Continued)

7.10 Approximate Lateral TransferFunctionsApproximate transfer functions that can be written out explicitly, and that reveal themain aerodynamic influences in a particular frequency range, can be very useful indesigning control systems. In Sec. 6.8 we presented two approximate second-ordersystems that simulate the complete fourth-order system insofar as the characteristicmodes are concerned. These same approximations can be used to get approximatetransfer functions for control response.

Page 256: Stability Control EtkinReid

248 Chapter 7. Response to Actuation of the Controls-Open Loop

10' \:=-------------\--;:---:-----:-:---=-,..J-\--\---------.....jDutch roll

approx.Spiral/roll

~~~~~ / approx.---------------

10°E----------i-----------+--~_\__-------iII

Dutch:roll :

IIIIII

10-1 '-:-_---J _ ____J..---J'-~....I_L...L.J __ ____J.._ __L___l.___l._L.Ll...w.' __ .....J..._....L---I........L~.L.J..LJ

10-2m (rad/s)

(a)

Spiral/roll~~~~~ / approx .

.......... _---- .•_----

II

Dutch:roll :

IIIIII

10-1 '--_----'_---"-----'----'~....L...JL...J...l __ ___L_ __'____L___L_'_..L...J....J.U.'__ ---"-_--'----"----"-""'-.J-L...L.I10-2 10-1 100 101

m (rad/s)(a)

Figure 7.27 Frequency-response functions, aileron angle input. Jet transport cruising at highaltitude. (a) Sideslip amplitude. (b) Sideslip phase. (c) Roll amplitude. (d) Roll phase. (e) Yaw-rateamplitude. if) Yaw-rate phase.

SPIRALIROLL APPROXIMATION •When aerodynamic control terms are added to (6.8,9) and the Laplace transform istaken, the result is

(7.10,1)

••,

Page 257: Stability Control EtkinReid

7.10 Approximate Lateral Transfer Functions 249

Spiral/rollapprox.

Q) (rad/s)(e)

o

----- Exact~ ,',______ I'

I I, ,SPiral/r~

~approx.

Dutch:roll' ,,,,

II I I I I I I I I I I I I I I I! I I I I II I I

180

120

60

-6

-12

-1810-2

Q) (rad/s)

(d)

Figure 7.27 (Continued)

In (7.10,1) the ay and j{ derivatives are as defined in Sec. 6.8, and the ~6 derivativesare

(7.10,2)

•where 8 is either 8a or B; From (7.10,1) we get the desired transfer functions. The de-nominators are all the same, obtained from (6.8,11) as

cSl + Ds + E (7.10,3)

Page 258: Stability Control EtkinReid

250 Chapter 7. Response to Actuation of the Controls-Open Loop

100c::-----------,------------,--- --,

Dutchrollapprox.

ro (rad/s)

(e)

60

-------I""--- --.....

~~

/ Spiral/rollapprox. ~~ ~~~_~_ approx.-----------------

1\

Dutch: ~"'<-- Exactroll'IIIII

I I I I I I II I I I I I I I I I I I I~~ I

\

180

120

-6

-12

-1810-2

ro (rad/s)

If)

Figure 7.27 (Continued)

and the numerators are (again using 8 for either 8a or 8r):

NU8 = a3s3 + a2s2 + als + ao

Nt/>8 = bIs + bo

Nr8 = d2s2 + d.s + do

Np8 = SNt/>8

(7.10,4) ..

Page 259: Stability Control EtkinReid

7.10 Approximate Lateral Transfer Functions 251

The coefficients in these relations are:

a3 = CiYa; az = -CiYa<9!,p + .N'r) - uo.N'aa

1= CiYa(9!,p.N'r - 9!,r.N'p) - uo(9!,a.N'p - 9!,p.N' a) + 9!,ag

ao = g(9!,r.N'a - 9!,a.N'r) (7.10,5)

b1 = CiYa9!,,,; bo = uo(9!,a.N'" - 9!,,,.N'a) + CiYa<9!,,,.N'r - 9!,r.N',,)dz = CiYa.N',,; d, = CiYa<9!,,,.N'p - 9!,p.N',,); do = g(9!,a.N'" - 9!,,,.N'a)

DUTCH ROLL APPROXIMATION

Following the analysis of Sec. 6.8 and adding control terms to the aerodynamics thereduced system equations are

••

iJ = CiY"v - uor + t::..CiYc

t = .N'"v + .N'7 + t::...N'c

From (7.10,6) we derive the canonical equation

x=Ax+Bcwhere x = [v r]T; c = [5a 5r]T

(7.10,6)

B = [0 CiYar].N's, .N'8

r

(7.10,7)

and where

N.N'8r = I~L8r + I~r (7.10,8)

z

With the system matrices given by (7.10,7) the approximate transfer functions arefound in the form of (7.2,8) (see Exercise 7.7) with

f(s) = SZ - (CiY" + .N'r)s + (CiY".N'r + uo.N',,) (7.10,9)

and

N"8a = -"o.N' 8aNraa = .N'aaS - CiY".N'aaN"8r = CiYa,S - (CiYar.N'r + UO.N'8)Nrar = .N'8rS - (CiY".N'8r - CiY8r.N',,) (7.10,10)

The accuracy of the preceding approximations is illustrated for the example jet trans-port on Figs. 7.26 and 7.27. Two general observations can be made: (1) The DutchRoll approximation is exact in the limit of high frequency, and (2) the spiral/roll ap-proximation is exact as w ---+ 0. In this respect the situation is entirely analogous tothat of the longitudinal case, with the spiral/roll corresponding to the phugoid, andthe Dutch Roll to the short-period mode. There are ranges of frequency in the middlewhere neither approximation is good. We repeat that lateral approximations must beused with caution, and that only the exact equations can be relied on to give accurateresults.

Page 260: Stability Control EtkinReid

252 Chopter 7. Response to Actuation of the Controls-Open Loop

7.11 Transient Response to Aileron and RudderWe have seen that useful lateral steady states are produced only by certain definitecombinations of the control deflections. It is evident then that our interest in the re-sponse to a single lateral control should be focused primarily on the initial behavior.The equations of motion provide some insight on this question directly. Following astep input of one of the two controls the state variables at t = 0+ are all still zero, andfrom (4.9,19) we can deduce that their initial rates of change are related to the controlangles by

iJ = ayarSrP = :f8}a + :faA (7.11,1)f = .N'a.Sa + .N'aA

The initial sideslip rate iJ is thus seen to be governed solely by the rudder and,since ay s; > 0, is seen to be positive (slip to the right) when 8r is positive (left rud-der). Of somewhat more interest is the rotation generated. The initial angular acceler-ation is the vector

cd = ip + kf (7.11,2)

The direction of this vector is the initial axis of rotation, and this is of interest. It liesin the xz plane, the plane of symmetry of the airplane, as illustrated in Fig. 7.28a. Theangle g it makes with the x axis is, of course,

ig = tan-I -=-

p

Let us consider the case of "pure" controls, that is, those with no aerodynamic cross-coupling, so that L8p = N8• = O. The ailerons then produce pure rolling moment and

(7.11,3)

{Principal axes

Zp- -

Zpr-- ....-I~~ ~

IE VBody axes

z(6)

Figure 7.28 Initial response to lateral control. (a) General. (b) Example jet transport.

Page 261: Stability Control EtkinReid

7.11 Transient Response to Aileron and Rudder 253

~114110106102989490 €o0

-4

-8

-12~A

,/"~ '" € '

Figure 7.29 Angle of axis of rotation.

the rudder produces pure yawing moment. In that case we get for On = 0 the angle gRfor response to rudder from

r r j{ s N li II' z 1tangR = p = p' = :£ '= I' 'l\.T = 1'" = Ix/lzx

li, zx•• lir z' zx(7.11,4)

and similarly for response to aileron:

tan gA = I zxlIz (7.11,5)

The angles gA' gR are seen to depend very much on the product of inertia Izx. When itis zero, the result is as intuitively expected, the rotation that develops is about eitherthe x axis (aileron deflected) or the z axis (rudder deflected). For a vehicle such as thejet transport of previous examples, with Ixp = OAlzp' the values of L; Iz, Izx given by(4.5,11) yield the results shown in Fig. 7.29. The relations are also shown to scale inFig. 7.28b for E = 20° (high angle of attack). It can be seen that there is a tendencyfor the vehicle to rotate about the principal x axis, rather than about the axis of theaerodynamic moment. This is simply because IxlIz is appreciably less than unity. Nowthe jet transport of our example is by no means "slender," in that it is of large spanand has wing-mounted engines. For an SST or a slender missile, the trend shown ismuch accentuated, until in the limit as aspect ratio ~ 0, both tan gR and tan gA tend totan E, and the vehicle rotates initially about the xp axis no matter what control isused!

SOLUTION FOR LARGE ANGLES

The preceding analysis shows how a lateral response starts, but not how it continues.For that we need solutions to the governing differential equations. As remarked ear-

Page 262: Stability Control EtkinReid

254 Chapter 7. Response to Actuation of the Controls-Open Loop

lier, at the beginning of this chapter, control responses can rapidly build up large val-ues of some variables, invalidating the linear equations that we have used so far.There is a compromise available that includes only some nonlinear effects that is use-ful for transport and general aviation airplanes, which are not subjected to violentmaneuvers. The compromise is to retain a linear representation of the inertia andaerodynamic effects, but to put in an exact representation of the gravity forces. Thisallows the angles cP, (J, and l/J to take on any values. As we shall see in the followingexample, the solution obtained is then limited by the airplane speed growing beyondthe range of linear validity, that is, it is an aerodynamic nonlinearity that then con-trols the useful range of the solution. When the procedure that led to (4.9,18 and 19)is repeated without the small angle approximations we get the following for 0

0= 0

(see Exercise 7.8):

t,qcos cP - rsin cP

sx--gsinOmu

w flZ- - g(l - cos o cos cP) + uoqmq

iJ

IlL- +1' A1\TI' zx'-"l'

x

fly- + gcos o sin cP - uorm

pt =

flN- + I' ATI' ~z

p + (q sin cP + r cos cP)tan (J

(q sin cP + r cos cP)sec0

(7.11,6)

(7.11,7)

The data for the B747 jet transport previously used was incorporated into the preced-ing equations. A step aileron input of -150 was applied at time zero, the other con-trols being kept fixed, and the solution was calculated using a fourth-order Runge-Kutta algorithm. The results are shown in Fig. 7.30. The main feature is the rapidacquisition of roll rate, shown in Fig. 7.30b, and its integration into a steadily grow-ing angle of bank (Fig. 7.30c) that reaches almost 900 in half a minute. Sideslip, yawrate, and yaw angle all remain small throughout the time span shown. As the airplanerolls, with its lift remaining approximately equal to its weight, the vertical componentof aerodynamic force rapidly diminishes, and a downward net force leads to negative(J and an increase in speed. After 30 seconds, the speed has increased by about10% of uo, and the linear aerodynamics becomes increasingly inaccurate. The maxi-mum rotation rate is p = .05 rad/s, which corresponds to p = 0.01. This is smallenough that the neglect of the nonlinear inertia terms in the equations of motion isjustified.

Page 263: Stability Control EtkinReid

o ----- ••UUl:l:l:~.~.~.~.:':::~:::: :::•.:- .•.••....-- ...•.•.

-20 •...•............

.e. 60

.~oo~ 40

0.09

0.08

.!!? 0.07"C~i- 0.06Ti'* 0.05>1!i 0.04'"Cl

~ 0.03

0.02

0.01

100.-------------------------,

80

20

°0!-...::;--5'-....;.;.-aj1~O=::;:;;;;..~1~5••••.•_~~ ••••••~~ ••••.•_~

Time, s(a) Velocity components

0.1.-------------------------,

5 10 15 20Time, s

(b) Angular velocity components

25 30

80

60

40

(c) Attitude angles

Figure 7.30 Response of jet transport to aileron angle; 13a = -15°. (a) Velocity components. (b)Angular velocity components. (c) Attitude angles.

255

Page 264: Stability Control EtkinReid

256 Chapter 7. Response to Actuation of the Controls-Open Loop

7.12 Inertial Coupling in Rapid ManeuversWe saw in the last section how to include nonlinear gravity effects in control re-sponse and how such effects manifest themselves in the response of a relatively se-date vehicle. Of the other two categories of nonlinearity-aerodynamic and inertial-little in a general way can be said about the first. Aerodynamic characteristics,especially for flexible vehicles at high subsonic Mach numbers, are too varied andcomplex to admit of useful generalizations. A very elaborate (and very costly!) aero-dynamic model is required for full and accurate simulation or computation. Not so,however, for the second category of nonlinearity. There is a class of problems, allgenerically connected, known by names such as roll resonance, spin-yaw coupling,inertia coupling, and so on (Heppe and Celinker, 1957; Phillips, 1948; Pinsker, 1958)that pertain to large-angle motions, or even violent instabilities, that can occur onmissiles, launch vehicles, and slender aircraft performing rapid rolling maneuvers.These have their source in the pq and pr terms that occur in the pitching and yawingmoment equations. A detailed analysis of these motions would take us beyond thescope of this text. Some is given in Etkin (1972), and much more is given in the citedreferences. One very important conclusion, due to Phillips (1948), is that there is aband of roll rates for airplanes within which the airplane is unstable. At lower rollrates, the usual stability criteria apply. At rates above the band the airplane is gyrosta-bilized in the way a spinning shell or top is. The lower of the critical roll rates for anormally stable airplane is given approximately by the lesser of

p2=_~ or ~I, - Ix Iy - I,

If the roll rate in a maneuver approaches or exceeds this value the possibility of adangerous instability exists.

7.13 Exercises7.1 A = [aij] is a (3X3) matrix. Demonstrate the statement made in the text with respect

to the numerator of (7.2,7) by writing out in full the adjoint of (sI - A).

7.2 Use the convolution theorem (Appendix A.3) to obtain an alternative proof of the the-orem for frequency response. That is, for a system with transfer function G(s) and in-put eiwt the response for t _ 00 is G(iw)eiwt

7.3 An airplane is flying at the speed V* for which the thrust curve is tangent to the dragcurve (Fig. 7.1). The throttle is then suddenly advanced to produce a higher thrustcurve, such as is seen in the figure. The pilot controls the elevator so as to maintainexactly horizontal flight, in which case the drag curve is as in the figure. The ultimatesteady state is at either P or Q.What will govern which it will be?

7.4 In a test flight procedure, the airplane is brought to a condition of steady horizontalflight in quiet air. The elevator is then displaced rapidly through a small angle, heldbriefly, and then returned as rapidly to its original position. Assume that the resultinginput can be treated as an impulse at t = 0 (see Sec. 7.3).

Page 265: Stability Control EtkinReid

7.13 Exercises 257

(a) Use the short period approximation (7.7,llb) to the transfer function for (J to de-rive a time domain solution for (J(t). Express the solution in terms of n, W, ho, andhi'

(b) Assuming that (J and t can be determined very accurately from the flight test data,and hence that n and W can be determined precisely, suggest how the experimen-tal data could be used to determine ho, hi' Co' and CI• Note that if Qo and QI couldlikewise be determined accurately, then the six equations (7.7,12) could in princi-ple be used to solve for the six aerodynamic derivatives on the right side of theequations.

7.5 Use the nondimensional derivatives of Table 7.2 to calculate the coefficients of thematrix equation (7.8,4).

7.6 (a) Reformulate the equations for the steady sideslip (7.8,3) to use cf> as input and (v,8a, 8r) as outputs.

(b) Derive (7.8,15)

7.7 Derive (7.10,10)

7.8 Derive (7.11,6) and (7.11,7).

7.9 An additional vertical control surface (8s) is added above the fuselage of an airplane,near the CG. It is capable of providing a side force, accompanied by a rolling mo-ment, given by

.:1Y = Yas8s; f1L = LaA

What condition must be satisfied if (8a, 8ro 8s) are to generate specified (Y, L, N)?

7.10 (a) Using the numerical data for the B747 example (Sec. 7.9), calculate the staticgains for each of the eight responses that correspond to (7.9,5)-that is, the val-ues of lG(iw);il for ta = O.

(b) Calculate the slopes of the high-frequency asymptotes for each of the eight fre-quency response amplitudes (express result in decades/decade).

(c) Assume that w = 0, that is, that a steady state exists in response to one of thecontrols being deflected, such that cf> = 15°. For each of the two controls-aileron and rudder---calculate the control angle, the sideslip angle, and the yawrate r.

7.11 The elevator of the B747 airplane is oscillated at a frequency a little below that of theshort-period mode.

(a) Use the results given in Fig. 7.18 to estimate the amplitude of the load factor ifthe elevator amplitude is 2°.

(b) What elevator amplitude would lift a passenger seated near the CG from the seat?(c) What elevator amplitude would cause the load factor to reach the FAR Part 25

limit maneuvering value of 2.5?

Page 266: Stability Control EtkinReid

258 Chapter 7. Response to Actuation of the Controls-Open Loop

7.14 Additional Symbols Introduced in Chapter 7

d(t) response to a unit step inputh(t) response to a unit pulse inputK static gain

:£a Lall: + I'zxNaM magnification factor

j{ a N JI~ + I'veLa6Ya YJm8(t) Dirac's delta function

<p phase angle

Page 267: Stability Control EtkinReid

CHAPTER 8

Closed Loop Control

8.1 General RemarksThe development of closed loop control has been one of the major technologicalachievements of the twentieth century. This technology is a vital ingredient in count-less industrial, commercial, and even domestic products. It is a central feature of air-craft, spacecraft, and all robotics. Perhaps the earliest known example of this kind ofcontrol is the fly-ball governor that James Watt used in his steam engine in 1784 toregulate the speed of the engine. This was followed by automatic control of torpedoesin the nineteenth century (Bollay, 1951), and later by the dramatic demonstration ofthe gyroscopic autopilot by Sperry in 1910, highly relevant in the present context.Still later, and the precursor to the development of a general theoretical approach,was the application of negative feedback to improve radio amplifiers in the 1930s.The art of automatic control was quite advanced by the time of the landmark four-teenth Wright Brothers lecture (Bollay, 1951)1. Most of what is now known as "clas-sical" control theory-the work of Routh, Nyquist, Bode, Evans, and others was de-scribed in that lecture. From that time on the marriage of control concepts withanalogue and digital computation led to explosive growth in the sophistication of thetechnology and the ubiquity of its applications.

Although open-loop responses of aircraft, of the kind studied in some depth inChap. 7, are very revealing in bringing out inherent vehicle dynamics, they do not inthemselves usually represent real operating conditions. Every phase of the flight ofan airplane can be regarded as the accomplishment of a set task-that is, flight on aspecified trajectory. That trajectory may simply be a straight horizontal line traversedat constant speed, or it may be a tum, a transition from one symmetric flight path toanother, a landing flare, following an ILS or navigation radio beacon, homing on amoving target, etc. All of these situations are characterized by a common feature,namely, the presence of a desired state, steady or transient, and of departures from itthat are designated as errors. These errors are of course a consequence of the un-steady nature of the real environment and of the imperfect nature of the physical sys-tem comprising the vehicle, its instruments, its controls, and its guidance system(whether human or automatic). The correction of errors implies a knowledge of them,that is, of error-measuring (or state-measuring) devices, and the consequent actuationof the controls in such a manner as to reduce them. This is the case whether control isby human or by automatic pilot. In the former case-the human pilot-the state in-formation sensed is a complicated blend of visual and motion cues, and instrumentreadings. The logic by which this information is converted into control action is only

'In 1951 most aeronautical engineers were using slide rules and had not heard of a transfer function!

259

Page 268: Stability Control EtkinReid

260 Chapter 8. Closed Loop Control

imperfectly understood, but our knowledge of the physiological "mechanism" that in-tervenes between logical output and control actuation is somewhat better. In the lattercase-the automatic control-the sensed information, the control logic, and the dy-namics of the control components are usually well known, so that system perfor-mance is in principle quite predictable. The process of using state information to gov-ern the control inputs is known as closing the loop, and the resulting system as aclosed-loop control or feedback control. The terms regulator and servomechanismdescribe particular applications of the feedback principle. Figure 8.1 shows a generalblock diagram describing the feedback situation in a flight control system. This dia-gram models a linear invariant system, which is of course an approximation to realnonlinear time-varying systems. The approximation is a very useful one, however,and is used extensively in the design and analysis of flight control systems. In the di-agram the arrows show the direction of information flow; the lowercase symbols arevectors (i.e. column matrices), all functions of time; and the uppercase symbols arematrices (in general rectangular). The vectors have the following meanings:

r: reference, input or command signal, dimensions (p Xl)z: feedback signal, dimensions (p Xl)e: error, or actuating, signal, dimensions (p X1)c: control signal, dimensions (mX 1)g: gust vector (describing atmospheric disturbances), dimensions (IX 1)x: airplane state vector, dimensions (nX 1)y: output vector, dimensions (qX1)

n: sensor noise vector, dimensions (q Xl)

Of the above, x and c are the same state and control vectors used in previouschapters. r is the system input, which might come from the pilot's control~er, from anexternal navigation or fire control system. or from some other source. It IS the com-mand that the airplane is required to follow. The signal e drives the system to make z

z

H(s)

Figure 8.1 A general linear invariant flight control system.

Page 269: Stability Control EtkinReid

8.1 GeneralRemarks 261

follow r. It will be zero when z = r. The makeup of the output vector y is arbitrary,constructed to suit the requirements of the particular control objective. It could be assimple as just one element of x, for example. The feedback signal z is also at the dis-cretion of the designer via the choice of feedback transfer function H(s). The choicesmade for D(s), E(s) and H(s) collectively determine how much the feedback signaldiffers from the state. With certain choices z can be made to be simply a subset of x,and it is then the state that is commanded to follow r.

The vector g describes the local state of motion of the atmosphere. This statemay consist of either or both discrete gusts and random turbulence. It is three-dimen-sional and varies both in space and time. Its description is inevitably complex, and togo into it in depth here would take us beyond the scope of this text. For a more com-plete discussion of g and its closely coupled companion G' the student should con-sult Etkin (1972) and Etkin (1981).

In real physical systems the state has to be measured by devices (sensors) suchas, for example, gyroscopes and Pitot tubes, which are inevitably imperfect. This im-perfection is commonly modeled by the noise vector n, usually treated as a randomfunction of time.

The equations that correspond to the diagram are (recall that overbars representLaplace transforms):

e=i'-z (a)(b)

(e) (8.1,1)(d)(e)

(8.1,2)

(8.1,3)

c = J(s)e

x = G(s)c + G'(s)gy=Dx+Ecz = H(s)(y + ii)

In the time domain (8.1,le) appears as

t = Ax + Be + Tg

It follows that

IG(s) = (sl - A)-IB and G'(s) = (sl - A)-IT

The feedback matrix H(s) represents any analytical operations performed on the out-put signal. The transfer function matrix J(s) represents any operations performed onthe error signal e, as well as the servo actuators that drive the aerodynamic controlsurfaces, including the inertial and aerodynamic forces (hinge moments) that act onthem. The servo actuators might be hydraulic jacks, electric motors, or other devices.This matrix will be a significant element of the system whenever there are power-assisted controls or when the aircraft has a fly-by-wire or fly-by-lightAFCS.

From (8.1,1) we can derive expressions for the three main transfer function ma-trices. By eliminating x, e, c, and z we get

[I + (00 + E)JH]y = (DG+ E)Jr - (DG+ E)JHn + DG'g (8.1,4)

from which the desired transfer functions are

Gyr = [I + (DG+ E)JHrl(DG + E)JGyn = -[I + (DG+ E)JHrl(DG + E)JHGyg = [I + (DG+ E)JHrIDG'

(a)(b) (8.1,5)

(e)

Page 270: Stability Control EtkinReid

262 Chapter 8. Closed Loop Control

The matrices that appear in (8.1,5) have the following dimensions:

D(q X n); G(n X m); E(q X m); J(m X p); H(p X q); G'(n X l)

The forward-path transfer function, from e to y, is

F(s) = (DG + E)J; dimensions (q Xp)

so the preceding transfer functions can be rewritten as

Gyr = (I + FH)-lF (a)

Gyn = -(I + FH)-lFH (b) (8.1,6)

Gyr = (I + FH)-lDG' (c)

Note that F and H are both scalars for a single-input, single-output system.When the linear system model is being formulated in state space, instead of in

Laplace transforms, then one procedure that can be used (see Sec. 8.8) is to generatean augmented form of (8.1,2). In general this is done by writing time domain equa-tions for J and H, adding new variables to x, and augmenting the matrices A and Baccordingly. An alternative technique for using differential equations is illustrated inSec. 8.5. There is a major advantage to formulating the system model as a set of dif-ferential equations. Not only can they be used to determine transfer functions, butwhen they are integrated numerically it is possible, indeed frequently easy, to add awide variety of nonlinearities. These include second degree inertia terms, dead bandsand control limits (see Sec. 8.5), Coulomb friction, and nonlinear aerodynamicsgiven as analytic functions or as lookup tables.

AN AERODYNAMICS VIEWPOINT

It is frequently helpful to view a feedback loop as simply a method of altering one ofthe airplane's inherent stability derivatives. When one of the main damping deriva-tives, Lp, Mq, or Nr, is too small, or when one of the two main stiffnesses MOl or Nf3 isnot of the magnitude desired, they can be synthetically altered by feedback of the ap-propriate control. Specifically let x be any nondimensional state variable, and let acontrol surface be displaced in response to this variable according to the law

t18 = kt1x; k = const

(Here k is a simplified representation of all the sensor and control system dynamics!)Then a typical aerodynamic force or moment coefficient Ca will be incremented by

(8.1,7)

This is the same as adding a synthetic increment

t1Cax = kCa/J (8.1,8)

to the aerodynamic derivative Cax' Thus if x be yaw rate and 8 be rudder angle, thenthe synthetic increment in the yaw-damping derivative is

t1Cnr = ic.; (8.1,9)

which might be the kind of change required to correct a lateral dynamics problem.This example is in fact the basis of the often-applied "yaw damper," a stability-aug-

Page 271: Stability Control EtkinReid

8.1 GeneralRemarks 263

mentation feature. Again, if x be the roll angle and {3the aileron, we get the entirelynew derivative

C1.p = kC1a" (8.1,10)

the presence of which can profoundly change the lateral characteristics (see Exercise8.1).

SENSORSWe have already alluded to the general nature of feedback control, and the need toprovide sensors that ascertain the state of the vehicle. When human pilots are in con-trol, their eyes and kinesthetic senses, aided by the standard flight information dis-played by their instruments, provide this information. (In addition, of course, theirbrains supply the logical and computational operations needed, and their neuro-mus-cular systems all or part of the actuation.) In the absence of human control, when thevehicle is under the command of an autopilot, the sensors must, of course, be physi-cal devices. As already mentioned, some of the state information needed is measuredby the standard flight instruments-air speed, altitude, rate of climb, heading, etc.This information mayor may not be of a quality and in a form suitable for incorpora-tion into an automatic control system. In any event it is not generally enough. Whenboth guidance and attitude-stabilization needs are considered, the state informationneeded may include:

Position and velocity vectors relative to a suitable reference frame.Vehicle attitude (e, <p, r/J).

Rotation rates (p, q, r).Aerodynamic angles (a, 13).Acceleration components of a reference point in the vehicle.

The above is not an exhaustive list. A wide variety of devices are in use to measurethese variables, from Pitot-static tubes to sophisticated inertial-guidance platforms.Gyroscopes, accelerometers, magnetic and gyro compasses, angle of attack andsideslip vanes, and other devices all find applications as sensors. The most commonform of sensor output is an electrical signal, but fluidic devices have also been used.Although in the following examples we tend to assume that the desired variable canbe measured independently, linearly, and without time lag, this is of course an ideal-ization that is only approached but never reached in practice. Every sensing device,together with its associated transducer and amplifier, is itself a dynamic system withcharacteristic frequency response, noise, nonlinearity, and cross-coupling. These at-tributes cannot finally be ignored in the design of real systems, although one can use-fully do so in preliminary work. As an example of cross-coupling effects, considerthe sideslip sensor assumed to be available in the gust alleviation system of Sec. 8.9.Assume, as might well be the case, that it consists of a sideslip vane mounted on aboom projecting forward from the nose. Such a device would in general respond notonly to 13 but also to atmospheric turbulence (side gusts), to roll and yaw rates, and tolateral acceleration ay at the vane hinge. Thus the output signal would in fact be acomplicated mathematical function of several state variables, representing several

Page 272: Stability Control EtkinReid

264 Chapter 8. Closed Loop Control

feedback loops. The objective in sensor design is, of course, to minimize all the un-wanted extraneous effects, and to provide sufficiently high frequency response andlow noise in the sensing system.

This brief discussion serves only to draw attention to the important design andanalytical problems related to sensors, and to point out that their real characteristics,as opposed to their idealizations, need finally to be taken into account in design.

8.2 Stability of Closed Loop SystemsFor linear invariant systems such as we have discussed above, the methods availablefor assessing stability include those used with open loop systems. One such methodis to formulate the governing differential equations, find the characteristic equation ofthe system, and solve for its roots. Another is to find the transfer function from inputto output and determine its poles. With the powerful computing methods available, itis feasible to plot loci of the roots (or poles) as one or more of the significant designparameters are varied, as we shall see in examples to follow. For the multivariablehigh-order systems that commonly occur in aerospace practise this is a very usefultechnique.

Let us now consider the stability of the loop associated with one particular in-put/output pair in the light of (8.1,6a). Since p = 1 and F and H are scalars, the trans-fer function is

GyrCs) = )U( )1 + F(s ri: S

F(s)(8.2,1)

The transfer functions F(s) and H(s) are ratios of polynomials in s, that is, F(s) =N/Dt, and H(s) = NzlDz. Equation (8.2,1) then leads to

NtDz (8.2,2)

The characteristic equation is evidently

Dt(s)Dz(s) + Nt(s)Nz(s) = 0 (8.2,3)

This should be contrasted with the characteristic equation for the airframe alone,which is D(s) = 0, where D is the denominator of G(s). The block diagram corre-sponding to (8.2,1) is shown in Fig. 8.2. FH is the open loop transfer function, that is,the ratio of feedback to error, fie. Its absolute value \FH\ is the open loop gain.

The stability of the system can be assessed from the frequency response F(iw)H(iw).It is clear that if there is a frequency and open loop gain for which FH = -1then un-

y

z

Figure 8.2 Consolidated block diagram of feedback controller.

Page 273: Stability Control EtkinReid

8.2 Stability of Closed Loop Systems 265

1m,...-----------------11

K = 8; Unstable

--2 -1 0

Figure 8.3 Nyquist diagram.

der those conditions the denominator of (8.2,1) is zero and Gyr(iw) is infinite. Whenthese conditions hold, the feedback signal that is returned to the junction point is pre-cisely the negative of the error signal that generated it. This means that the systemcan oscillate at this frequency without any input. This is exactly the situation with thewhistling public address system. For then the acoustic signal that returns to the mi-crophone from the loudspeakers, in response to an input pulse, is equal in strength tothe originating pulse. Clearly the point (-1,0) of the complex plane has special sig-nificance. Nyquist (1932) has shown how the relationship of the frequency responsecurve (the Nyquist diagram) of FH to this special point indicates stability (see Fig.8.3). In brief, if the loop gain is <1 when the phase angle is 180°, or if the phase is<180° when the gain is unity, then the system is stable. The amounts by which thecurve misses the critical point define two measures of stability, the gain margin andphase margin, illustrated on the Nichols diagram (see McLean, 1990) of Fig. 8.4. The

10,--------------.,-----..----------,

Q)"0:J

'§, 1.0t:------------rt-+-----------j'"~

Phasemargin

-0.1~---~ _L _ __L._--l L_ ..L_ ....J

-360

Gainmargin

-300 -240 -180 -120Phase, degrees

Figure 8.4 Nichols diagram. K = 2.

-60 o

Page 274: Stability Control EtkinReid

266 Chapter 8. Closed Loop Control

examples of Figs. 8.3 and 8.4 are for the open loop transfer function

FH = K(s + 0.125)s2(0.15s2 + 0.8s + 1)

8.3 Phugoid Suppression: Pitch Attitude ControllerThe characteristic lightly damped, low-frequency oscillation in speed, pitch attitude,and altitude that was identified in Chap. 6, was seen in Chap. 7 to lead to large peaksin the frequency-response curves (Figs. 7.14 to 7.18) and long transients (Fig. 7.20).Similarly, in the control-fixed case, there are large undamped responses in this modeto disturbances such as atmospheric turbulence. These variations in speed, height,and attitude are in fact not in evidence in actual flight; the pilot (human or automatic)effectively suppresses them, maintaining flight at more or less constant speed andheight. The logic by which this process of suppression takes place is not unique. Inprinciple it can be achieved by using feedback signals derived from anyone or acombination of pitch attitude e, altitude h, speed v, and their derivatives. In practice,the availability and accuracy of the state information determines what feedback isused.

Since the phugoid oscillation cannot occur if the pitch angle (J is not allowed tochange (except when commanded to), a pitch-attitude-hold feature in the autopilotwould be expected to suppress the phugoid. This feature is commonly present in air-plane autopilots. We shall therefore look at the design of an attitude hold system forthe jet transport of our previous examples. Pitch attitude is readily available from ei-ther the real horizon (human pilot) or the vertical gyro (autopilot). Consider the con-troller illustrated in Fig. 8.5. From (8.2,1) we see that the overall transfer function is

jj GIJB.(s)J(s)-=- = (8.3,1)Be 1 + GIJB.(s)J(s)

If we write GIJB.(s) = N(s)/D(s), and J(s) = N'(s)!D'(s), then the characteristic equa-tion is

D(s)D'(s) + N(s)N'(s) = 0 (8.3,2)

To proceed further we need explicit expressions for the above transfer functions.Since B is an important variable in both the short period and phugoid modes, it mightbe expected that neither of the two approximate transfer functions for G IJB.derived inSec. 7.7 would serve by itself. We therefore use the exact transfer function derived

controller aircraft

9

z

Figure 8.5 Pitch attitude controller.

Page 275: Stability Control EtkinReid

8.3 Phugoid Suppression: Pitch Attitude Controller 267

••

1.5r------------7-~-------,k2 = -0.5 Short-period

1.1 f-

k2 =0

0.3 f-

k2 =0

Phugoid \k2=-0.5~........-!:

o I-------;-----r-----O-.,.....--.....,;; H

.,g' 0.7 f-E

-0.1_1 -0.75 -0.50 -0.25 0

Real s

Figure 8.6 Root locus of pitch controller with proportional control.

from the full system of linearized longitudinal equations of motion. Then N(s) isgiven by (7.7,2) and D(s) by (6.2,2). The result is

-(1.158s2 + 0.3545s + 0.003873)GfJ6e = S4 + 0.750468s3 + 0.935494sz + 9.463025 X 1O-3s + 4.195875 X 10 3

(8.3,3)

As to J(s), a reasonable general form for this application is

k1J(s) = - + kz + k3s

S(8.3,4)

For obvious reasons, the three terms on the right hand side are called, respectively,integral control, proportional control and rate control, because of the way they oper-ate on the error e. The particular form of the controlled system, here GfJ6e(s), deter-mines which of k1, kz, k3 need to be nonzero, and what their magnitudes should be forgood performance. Integral control has the characteristic of a memory, and steady-state errors cannot persist when it is present. Rate control has the characteristic of an-ticipating the future values of the error and thus generates lead in the control actua-tion. In using (8.3,4), we have neglected the dynamics of the elevator servo actuatorand control surface, which would typically be approximated by the first-order trans-fer function 1/(1 + TS). Since the characteristic time of the servo actuator system, T,

is usually a small fraction of a second, and we are interested here in much longertimes, this is a reasonable approximation.

For the example airplane at the chosen flight condition it turns out that we needall three terms of (8.3,4) to get a good control design. This might not always be thecase. Let us first look at the use of proportional control only, in which case J is a con-stant gain, k2• To select its magnitude, we use a root locus plof of the system, Fig.8.6, in which the locus of the roots of the characteristic equation of the closed loopsystem are plotted for variable gain k2• We see that at a gain of about -0.5 thephugoid mode is nearly critically damped, that is, it is about to split into two realroots. At this gain, the phugoid oscillation is effectively eliminated. We note that at

2The term root locus is used throughout this chapter with the meaning ordinarily ascribed to it in thecontrol theory literature.

Page 276: Stability Control EtkinReid

268 Chapter 8. Closed Loop Control

the same gain the short period roots have moved in the direction of lower damping.The response of the aircraft to a unit step command in pitch angle with only propor-tional control is shown in Fig. 8.7a. It is clear that this is not an acceptable response.There is a large steady-state error (steady-state error is a feature of proportional con-trol) and the short-period oscillation leads to excessive hunting. The steady-state er-ror could be reduced by increasing k2 (see Exercise 8.2), but this would further de-crease the short period damping. \

2

J= -0.5 )c~v~ r----r---

20 40 60Time, S

(a)

80 100 120

1.5

e 1

0.5

2,.....-------,---------r-----_J = -0.5(1+ 1/s)

1.5hl\------+------+---------j

0.51+-------+------+-------1

20 40 60Time,s

(b)

2

J = -0.5(s + 1 + 1M

./

//

10 20 30

1.5

e 1

0.5

o oTime,s

(e)

Figure 8.7 Response of pitch angle to unit step command. (a) With proportional control. (b) Withproportional plus integral control. (c) With proportional, integral, and rate control.

Page 277: Stability Control EtkinReid

8.3 Phugoid Suppression: Pitch Attitude Controller 269

We digress briefly to explore the reason for the damping behavior. It was notedpreviously (Sec. 6.8 and Exercise 6.4) that the term of next-to-highest degree in thecharacteristic equation gives "the sum of the dampings." That is, the coefficient of S3

in (8.3,3) is the sum of the real parts of the short-period and phugoid roots. Nowwhen J = k2 the closed loop characteristic equation (8.3,2) becomes D(s) + k2N(s).That is we add a second degree numerator to a fourth degree denominator, leavingthe coefficient of S3 unchanged. Thus any increase in the phugoid damping can onlycome at the expense of that of the short-period mode. This is exactly what is seen inFig. 8.6. The shifts of the two roots in the real direction are equal and opposite.

To eliminate the steady-state error, we use integral control and choose

J = -0.5 (1 + ~)

The result is shown in Fig. 8.7b. The steady-state error has been eliminated, but theshort-period oscillation is now even less damped. Now the damping of the short-period mode is governed principally by Mq [see (6.3,14)], so in order to improve itwe should provide a synthetic increase to Mq• A signal proportional to q is readily ob-tained from a pitch-rate gyro. Since q = iJ in the system model we are using, we ac-complish this by adding a third term to J:

(8.3,5)

J = -0.5 (s + 1 + ~)

The result, shown in Fig. 8.7c is an acceptable controller, with little overshoot andno steady-state error. A commanded pitch attitude change is accomplished in about10 sec. Note that in this illustration, all the constants in (8.3,4) are the same, that is,-0.5. Fine-tuning of these could be used to modify the behavior to reduce the over-shoot or speed up the response. Throughout this maneuver, the elevator angle remainsless than its steady-state value (which it approaches asymptotically), so that the gainsused are indeed much smaller than the elevator control is physically capable of pro-viding (see Exercise 8.2).

The preceding analysis does not reveal the underlying physics of why k2 dampsthe phugoid. This can be understood as follows. An angle () in the low frequencyphugoid implies vertical velocity (i.e., It = V(). Now a positive elevator angle pro-portional to a slowly changing () implies a negative increment in angle of attack andhence in the lift as well. Thus k2 leads to a vertical force (downward) 1800 out ofphase with the vertical velocity (upward), exactly what is required for damping.

Although the phugoid oscillation has been suppressed quite successfully by thestrategy employed above, it should be remarked that in the example case neither thespeed nor the altitude has been controlled. As a consequence, the speed drifts ratherslowly back to its original value, and the altitude to a new steady state.

Finally it should be noted that a controller design that is correct for one flightcondition, in this case high speed at high altitude, may not be acceptable at all speedsand altitudes, for example, landing approach. In the real world of APeS design, thisproblem leads, as in most engineering design, to compromises between conflictingrequirements. If the economics of the airplane justifies it, gain scheduling can beadopted; that is, the control gains are made to be functions of speed, altitude, andconfiguration.

(8.3,6)

Page 278: Stability Control EtkinReid

270 Chapter 8. Closed Loop Control

8.4 Speed ControllerThe phugoid makes its presence known not only in the form of transient perturba-tions from a steady state, but also in maneuvers, as illustrated in Sec. 7.7. We sawthere for example that in changing from level to climbing flight by opening the throt-tle (Fig. 7.21) there results a protracted, weakly damped approach to the new statethat would take more than 10 min to complete. Transitions from one value of 'Yto an-other are obviously not made in this manner, and the pilot suppresses the oscillationin this case as well. Provided that the correct (J is known for the climb condition, thesame technique as discussed above would work, that is, control operating on pitch-attitude error. We illustrate an alternative concept that does not require any knowl-edge of the final correct pitch attitude, but that uses speed error alone. It is not self-evident how speed should be controlled, in the light of the discussion in Sec. 7.1. Wesaw there that both elevator and throttle influence the speed, but that the short- andlong-term effects of each of these controls are quite different-the throttle principallyaffects the speed only in the short term. For a change of steady-state speed, the eleva-tor must be used. Clearly, a sophisticated speed control might use both. We shall seein this example, however, that when the primary aim is to suppress the phugoid,which is a very long period oscillation, the goal can be achieved with the elevatoralone. Figure 8.8 shows the system.

The command is the speed Uc and the feedback signal is the actual speed u. Foroutput we choose speed and flight-path angle, that is, Y = [u 'Yf. The control vectoris c = [5e 5pf of which only the elevator is in the feedback loop. Since the con-trolled variable is u, which does not change appreciably in the short-period mode, wecan use the phugoid approximation for the aircraft transfer function matrix G(s),which is the (2X2) matrix of transfer functions from c to y:

G(s) = [Gu8e GU8p] (8.4,1)G'Y8e G'Y8p

Two of the elements of G are implicit in (7.7,7), since G'Y8 = GII8 - Ga8 where 5stands for either 5e or 8p• The remaining two are (see Exercise 8.4)

X8p uoM.",sGU8p = - -;;;-f(s)

z; z,M --M-X W mUm

G -~------'Y8p - m f(s)

••

(8.4,2)

'Y

u

u

Figure 8.8 Speed controller.

Page 279: Stability Control EtkinReid

8.4 Speed Controller 271

Ultimately we shall want to calculate the time responses of u and 'Yto a throttleinput ~p' So the transfer functions we need are the two corresponding closed looptransfer functions. If we denote these by Gu8p(s) and G-ys/s), respectively, we findthat they are given in terms of the aircraft transfer functions by (see Exercise 8.4)

(8.4,3)

Each of the aircraft transfer functions in (8.4,3) can, as usual, be expressed as a ratioof two polynomials, for example:

and (8.4,4)

When this is done (8.4,3) becomes

(8.4,5)

(8.4,6)

We know that the denominator of a transfer function is the characteristic polynomial.We also know that a linear invariant system of the kind under discussion can haveonly one independent characteristic equation. Thus we have an apparent paradox,since the denominator of (8.4,6) is not the same as that of (8.4,5), having the extrafactor f(s). Now it can be shown (see Exercise 8.6) that f is a factor of the bracketedterm in the numerator of (8.4,6), and hence that it divides out of the right side andleaves the same characteristic polynomial as in (8.4,5).

As indicated above, the second-order phugoid approximation should be expectedto be reasonable for this case. We shall therefore use it to choose the gains in J(s), butat the end will check the solution for suitability with the exact fourth-order equations.To this end we examine the effect of J(s) on the characteristic equation, that is, on

(8.4,7)

f(s) is given by (6.3,9):

f(s) = As2 + Bs + C

Nus. is given by (7.7,7):

and for J(s) we use

(8.4,8)

so that Dj = 1 and N, = k1 + k2s. Note that the ~s term implies a signal proportionalto acceleration. Such a signal could be obtained from an x-axis accelerometer or by

Page 280: Stability Control EtkinReid

272 Chapter 8. Closed Loop Control

0.4

0.6 r--------------------

0.2

K, = 0.02

~ 0.1g> 0 !--+-------D--oQ--------__=__,E

-0.2

-0.4

-0.6_L1----'------o.L.-

6---_...L

O.-4---_...J.

0.-2------'0

Real s

Figure 8.9 Speed controller. Root locus plot of Gu8e• Phugoid approximation.

differentiating the signal from the speed sensor. The closed loop characteristic equa-tion then becomes:

A's2 + B' s + C' = 0 (a)

where A' = A + a1k2 (b) (8.4,9)

B' = B + a1k1 + aOk2 (c)

C' = C + aOk1 (d)

The numerical values of the constants for the example jet transport are

A = 2.721 X 107

B = 2.633 X 105

C = 1.376 X 105

al = 8.218 X 108

ao = 3.653 X 108

0.09

0.08

0.07NUl,g- 0.06"0<0'N 0.05

oO!

*' 0.04a::

0.1 .....----.,..---------------~---,

(8.4,10)

0.03

0.02

0.01o L------l_---.l._.....L_...L_...l..-_.L-_L----l_---I

-0.002 0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016Proportional. k, rad/fps

Figure 8.10 Speed controller gain relation for' = 1. Phugoid approximation.

Page 281: Stability Control EtkinReid

8.4 Speed ControUer 273

5.----------r--------r--------,

J = .005(35 + 1)Input: throttle, I)p '" - +

"'o~.Eo"ti'Q)Q)0.

U) -5f--------+-------+--------1

-100'---------::"20:----------:4-'-0-------='60

Time, 5

Figure 8.11 Speed controller-phugoid approximation. Speed response to throttle input.

To assess what range of values of k, and k2 would be appropriate, we use threeguides:

1. A reasonable elevator angle for, say, a 10 fps (3.048 mls) speed error2. The root locus plot for GuBe

3. The graph of "-2 vs. k1 for critical damping

(l) The first of these is arrived at by noting that 10 of elevator for 10 fps speed errorgives a k, of 0.0017 rad/fps. (2) The root locus plot is shown on Fig. 8.9 and indicatesthat the open loop roots can be moved very appreciably with a proportional gain aslow as 0.005. (3) For the third guide, we note that critical damping corresponds toB,2 - 4A'C' = O.With the aid of (8.4,9) and (8.4,10) this leads to an algebraic rela-tion between k1 and "-2 that is solved for the graph shown on Fig. 8.10. The usefulrange of gains is the space below the curve, which corresponds to damped oscilla-tions. The farther from the curve, the more overshoot would be expected in the re-sponse. We have for illustration arbitrarily chosen the gains indicated by the pointmarked on the graph, without regard for whether it is optimum. When used to calcu-late the response of airplane speed to application of a negative step in thrust, with thephugoid approximation, the result is as shown in Fig. 8.11. The throttle input corre-

5

"-/ J = 0.005(35 + 1)I) =_..1.p 6

"'0.Eo"ti'Q)Q)0.

U) -5

-10o 20 40 60Time, 5

Figure 8.12 Speed controller-s-exact equations. Speed response.

Page 282: Stability Control EtkinReid

274 Chapter 8. Closed Loop Control

O~--------------------,J = 0.005 (35 + 1)

II =_..lp 6

-0.025

/'Ysst -0.05 r--\--~.-.."""""""",,=-_--_-L----__ ....j

Eco(9

-0.075

-0.1 '-----'-----'--_--'- __ ---JL- __ ...l...- __ ....Jo 60

Time, 5

Figure 8.13 Speed controller-exact equations. Gamma response.

20 40 80 100

sponds to a steady-state descent angle of a little less than 3°. The maximum speed er-ror, which is seen to be less than 3 fps at an initial speed of 774 fps, would probablynot be perceptible to the pilot. This suggests that the chosen gains are probably nottoo small. The maximum elevator angle during this maneuver is less than 2° (see Fig.8.14) so the gains are not excessive either.

To assess the performance of the controller with certainty, it is necessary to usethe exact equations. The full matrix A for this example is (6.2,1), and B is (7.6,4).The most important elements of the solution are displayed in Figs. 8.12 to 8.14. Theresult for the speed in Fig. 8.12 confirms that the phugoid approximation is indeedgood enough for preliminary design. Figure 8.13 demonstrates that the steady-stateflight path angle is reached, with a small overshoot, in about 20 s. Figure 8.14demonstrates that the elevator angle required to achieve this is small. To understandthe physics of the maneuver, it is helpful to look at the angle of attack variation,graphed in Fig. 8.15. It shows that there is a negative "pulse" in a that lasts about10 s. This causes a corresponding negative pulse in lift, which is the force perpendic-ular to the flight path that is required to change its direction.

Finally, these graphs should be contrasted with those of Fig. 7.21, which showthe uncontrolled response to throttle. Feedback control has made a truly dramatic dif-ference!

0.03

~

J = 0.005 (35 + 1)II =_..1..p 6

-L....

0.05

"0~ 0.01~.B 0~ -0.01

rn

-0.03

-0.05 o 10 20 30Time, 5

Figure 8.14 Speed controller-exact equations. Elevator angle.

Page 283: Stability Control EtkinReid

8.5 Altitude and Glide Path Control 275

0.05,-----------------------,

J = 0.005(38+ 1)II =_..1.p 60.025

"0

f!! I -------7""~:::==========:==~ro O~s:a.<i:-0.025

-0.050L....-------1

LO

------2LO ---------"30

Time. 5

Figure 8.15 Speed controller--exact equations. Angle of attack response.

8.5 Altitude and Glide Path ControlOne of the most important problems in the control of flight path is that of following aprescribed line in space, as defined for example by a radio beacon, or when the air-plane flies down the ILS glide slope. We discuss this case by considering first a sim-ple approximate model that reveals the main features, and then examining a more re-alistic, and hence more complicated case.

FLIGHT AT EXACTLY CONSTANT HEIGHT-SPEED STABILITY

The first mathematical model we consider can be regarded as that corresponding tohorizontal flight when a "perfect" autopilot controls the angle of attack in such a wayas to keep the height error exactly zero. The result will show that the speed variationis stable at high speeds, but unstable at speeds below a critical value near the mini-mum drag speed. Neumark (1950) recounts that this criterion was first discovered in1910 by Painleve, and that it was at first accepted by aeronautical engineers and sci-entists, but later, on the basis of the theory of the phugoid, which showed no such ef-fect, was rejected as false. In fact, to the extent that pilots can control height error byelevator control alone, that is, to the extent that they approximate the ideal autopilotwe have postulated, the instability at low speed will be experienced in manual flight.Since speed variation is the most noticeable feature of this phenomenon, it is com-monly referred to as speed stability.

We could analyze this case by applying (4.9,18) to the stated flight condition.However, it is both simpler and more illuminating to proceed directly from first prin-ciples. The airplane is flying on a horizontal straight line at variable speed V. It is im-plied that a is made to vary, by controlling Be' in such a way that the lift is kept ex-actly equal to the weight at all times. The equation of motion is clearly

mV= T-D (8.5,1)

where T is the horizontal component of the thrust, and D is the drag. Since the speedcannot change very rapidly, then neither does a, and we can safely ignore any effectsof q and a on lift and drag. Consequently, T and D are simply the thrust and drag or-

Page 284: Stability Control EtkinReid

276 Chopter 8. Closed Loop Control

D,TD(L=W)

T (Canst. throttle)

L...--------=-----------;~VV*

Figure 8.16 Performance graph.

dinarily used in performance analysis, as displayed in Fig. 8.16. We denote the refer-ence thrust and drag by Toand Do and define the stability derivatives

Tv = aT/aV and Dy = aD/aVT - D = (To + TydV) - (Do + DydV)so that

Since V = Vo + dV, and To = Do, (8.5,1) becomes

mdV= (Ty - Dy)dV (8.5,2)

This first-order differential equation has the solution

dV= ae"with A = (Ty - Dy)/m (8.5,3)

Tv and Dy are the slopes of the tangents to the thrust and drag curves at their intersec-tion. If they intersect at a point such as P in Fig. 8.16, then Ty < Dy, A is negative,and the motion is stable. If, on the other hand, the flight condition is at a point suchas Q, the reverse is the case. A is then >0, and the motion is unstable. If when flyingat point Q there is an initial error in the speed, then it will either increase until itreaches the stable point P or it will decrease until the airplane stalls. The stable andunstable regimes are bounded by the speed V*, which is where the thrust curve is tan-gent to the drag curve. V* will be the same as Vmd of Fig. 7.1 if Ty = O.Hence theappellation "back side of the polar" is used to describe the range V < V*, with refer-ence to the portion of the aircraft polar (the graph of CL vs. CD) for which CL isgreater than that for maximum LID.

Although we have analyzed only the case of horizontal flight, the result is similarfor other straight-line flight paths, climbing or descending (see Exercise 8.8). Flightin the unstable regime can indeed occur when an airplane is in a low speed climb orlanding approach. This speed instability is therefore not entirely academic, but canpresent a real operational problem, depending on by what means and how tightly theaircraft is constrained to follow the prescribed flight path. An important point insofaras AFCS design is concerned is that for speeds less than V* it is not possible to lock

-,

Page 285: Stability Control EtkinReid

8.5 Altitude and Glide Path Control 277

exactly onto a straight-line flight path, and at the same time provide stability, usingthe elevator control alone, no matter how sophisticated the controller! To achieve sta-bility it is mandatory to use a second control. This would most commonly be thethrottle, but in principle spoilers that control the drag could also be used.

EXAMPLE-AN ALTITUDE CONTROLLER

In view of the above, we illustrate an altitude controller that also incorporates controlof speed, using once again our example jet airplane. This time we make the systemmodel more realistic by including first-order lag elements for the two controls: thatfor the elevator is mainly associated with its servo actuator (time constant 0.1 s); andthat for the throttle with the relatively long time lag inherent in the build up of thrustof a jet engine following a sudden movement of the throttle (time constant 3.5 s). An-other feature that is incorporated to add realism to the example is a thrust limiter. Be-cause transport aircraft inherently respond slowly to changes in thrust, the gains cho-sen to give satisfactory response for very small perturbations in speed will lead to ademand for thrust outside the engine envelope for larger speed errors. We have there-fore included a nonlinear feature that limits the thrust to the range 0 ::;;T ::;;1.1To.This contains the implicit assumptions (quite arbitrary) that the airplane, flying nearits ceiling, has 10% additional thrust available, and that idling engines correspond tozero thrust.

At the same time this example illustrates an alternative approach to generatingthe analytical model of the system, in terms of its differential equations. In the previ-ous illustrations we have, by contrast, used what may be termed "transfer function al-gebra" to arrive at transfer functions of interest, and then used these to obtain what-ever results were desired. The end result of the modeling to follow is a system ofdifferential equations that is then integrated to get time solutions. Since the limiter isinherently a nonlinear element, it is in any case not possible to include it in a transferfunction based analysis.

The system block diagram is shown in Fig. 8.17. The commanded speed and alti-tude are the reference values Uo and ho, so that the two corresponding error signalsare - au and - Sh. Note that h is the negative of ZE used in Chap. 4. The inner loopfor ()is that previously studied in Sec. 8.3, with the lis) modified to account for the

'."'..~ Cp(s) ~ =+= Allp Ys--+- Jp(s) ~

Controller Limiter Engine G (s)

he=ho +~ Be + ee All. Ik J.(s) I------ - Airframe

Controller& elevator

u

B

h

Figure 8.17 Altitude-hold controller.

Page 286: Stability Control EtkinReid

278 Chapter 8. Closed Loop Control

elevator servo actuator. The logic of the outer loop that controls h warrants explana-tion. If there is an initial error in h, say the altitude is too low, then in order to correctit, the airplane's flight path must be deflected upward. This requires an increase inangle of attack to produce an increase in lift. The angle of attack and the resulting liftcould of course be produced by using an angle of attack vane as sensor, and no doubtan angle of attack commanded to be a function of height error would be very effec-tive. It might be preferred, however, to use the vertical gyro as the source of the sig-nal, and since short-term changes in 0 are effectively changes in a, then much thesame result is obtained by using (J as the commanded variable. We have chosen to usestability axes, so that in the steady state, when I1h is zero, the correct value of 0 isalso zero. Thus, in summary, the system commands a pitch angle that is proportionalto height error and the inner loop uses the elevator to make the pitch angle follow thecommand. While all this is going on the speed will be changing because of both grav-ity and drag changes. The quickest and most straightforward way of controlling thespeed is with the throttle, and the third loop accomplishes that. (The symbols Y4 andYs denote the inputs to the limiter and the airframe, and are elements of the state vec-tor derived below.)

The Differential Equations

The basic matrix differential equation of the airframe, with ()o = 0, and I1zE = -l1his obtained from (4.9,18) and (7.6,4). Up to this point, we have neglected engine dy-namics and in effect regarded thrust as proportional to 5p• The matrix B of (7.6,4) isstructured in that way. To accommodate the facts that 5p actually represents the throt-tle setting, not the thrust, and that the two are dynamically connected, we need to in-troduce two new symbols, Ys and c*. The quantity Ys, when multiplied by XlJp' etc.yields the aerodynamic force and moment increments l1Xe, etc., and c* is defined be-low. The differential equation of the airframe is then

x = Ax + Bc*

where x = [l1u w q 0 I1hfc* = [5e ysfA = [aij]

B = [bij]

To obtain the differential equations of the three control elements, we begin withtheir transfer functions, which are specified for this example to be

Je(s) = (aoS-1 + al + a2s)(1 + TeS)-1

= (ao + als + a2s2)(s + TeS2)-1 (8.5,5)

Cis) = (boS-1 + bl + b2s)= (bo + b.s + b2s2)/S (8.5,6)

Jp(s) = 1/(1 + TpS) (8.5,7)

The first two of these contain proportional + rate + integral controls, all of whichwere found to be needed for good performance. The time domain equations that cor-respond to the elements of the controller are then as follows (verify this by takingtheir Laplace transforms):

(8.5,4)

Page 287: Stability Control EtkinReid

8.5 Altitude and Glide Path Control 279

()c = -M.h (a)

'Ted5e + ss, = aoe(J+ ale(J + a2ii(J (b) (8.5,8)

Y4 = boeu + bleu + b2iiu (c)'TpYs= d5p - Ys (d)

After substituting the expressions for the two error signals, (8.5,8) yield three equa-tions for the controls

'Ted5e + ss, = -ka2dh - a20 - kaldh - aJJ - kaodh - aoO (a)Y4 = -b2dU - bldit - bodu (b) (8.5,9)

'TpYs= d5p - Ys (c)

For convenient integration we want a system of first-order equations and thereforehave to do something about the second derivatives in (8.5,9). Since {J = q we can re-place 0 with CJ.For the other second derivatives, we introduce three new variables, asfollows:

YI = ditY2 = dhY3 = ss,

(a)

(b) (8.5,10)

(c)

With these definitions, (8.5,9a and b) can now be rewritten in terms of first deriva-tives as

'TeY3+ Y3 = -(a2CJ + ka2Y2 + a.q + aoO + kaodh + ka1Y2)Y4 = -(b2YI + bodu + bIYI)

The state vector now consists of the original five variables from (8.5,4) plus the twocontrol variables d5e and d5p, plus the five Yi defined above, making 12 in all. Wetherefore require 12 independent equations. From the foregoing equations (8.5,4)(8.5,10), (8.5,11), and (8.5,9c) we can get 11 of the required differential equations.That for YI is obtained from (8.5,lOa) by differentiating the first component of (8.5,4)and that for Y2 by differentiating the fifth. The result of that operation is

YI = a12w + allYl + al4q + bllY3 + bl2ySY2 = -w + uoq

Finally, the 11 independent differential equations are assembled as follows:

dit = YI

w = a21du + a22w + a23q + b21d5e

CJ= a31du + a32w + a33q + b31d5e

{J=q

dh = Y2ss, = Y3

Ys = (-Ys + d5p)/'Tp

YI = allYl + al2w + al4q + bllY3 + bl2ySY2 = uoq - WY3 = -(a2CJ + ka2Y2 + alq + aoO + kaodh + kalY2 + Y3)/'Te

Y4 = -(b2YI + bodu + bIYI)

(8.5,11)

(8.5,12)

(8.5,13)

Page 288: Stability Control EtkinReid

280 Chapter 8. Closed Loop Control

Solutions

The above equations contain first derivatives on the right side as well as on the leftside and hence are not in the canonical form. This is no impediment to numerical in-tegration, however, since if the derivatives are calculated in the sequence given, eachone that appears on the right has already been calculated in one of the precedingequations by the time it is needed. The twelfth and final relation needed is that whichdescribes the limiter, in the form

ASp = !(Y4)

From the values of CDo and CLo given in Sec. 6.2, we find that Do = To =0.0657W. In Sec. 7.6 it was given that Sp = 1 corresponds to a thrust of 0.3W. It fol-lows that zero thrust corresponds to Y4 = -0.0657/.3 = -0.219. The nonlinear rela-tionship for Y4 is therefore implemented in the computing program by a programfragment equivalent to

ASp = Y4IF Y4 < -0.219 THEN A8p = -0.219IFY4> 0.10 THEN ASp = 0.10

where the maximum engine thrust has been assumed to be 10% greater than cruisethrust. Equations (8.5,13 and 8.5,14) are convenient for numerical integration. Wehave calculated a solution using simple Euler integration of the equations for the ex-ample jet transport with the matrices A and B given in Sees. 6.2 and 7.6, and with thefollowing control parameters: 'Te = 0.1; 'Tp = 3.5; k = 0.0002; ao = al = a2 = -0.5;bo = 0.005; bl = 0.08; b2 = 0.16. Figure 8.18 shows the performance obtained in re-sponse to an initial height error of 500 ft.

It is seen that the height error is reduced to negligible proportions quickly, inabout 20 s, accompanied by a theta pulse of similar duration and peak magnitudeabout 7°. Even with extreme throttle action, the speed takes more than 2 min to re-cover its reference value. This length of time is inherent in the physics of the situa-tion and cannot be shortened significantly by changes in the controller design. On theother hand, there is no operational requirement for more rapid speed adjustmentwhen cruising at 40,000 ft.

The peak elevator angle needed is less than 3°, but the thrust drops quickly tozero, stays there for about 30 s, then increases rapidly to its maximum. Towardthe end of the maneuver the throttle behaves linearly and reduces the speed errorsmoothly to zero.

(8.5,14)

8.6 Lateral ControlThere are five lateral state variables that can be used readily as a source of feedbacksignals-{ v, p, r, ep, l/J}; v from a sideslip vane or other form of aerodynamic sensor,p and r from rate gyros, and ep, l/J from vertical and directional gyros. Lateral acceler-ation is also available from an accelerometer. These signals can be used to drive thetwo lateral controls, aileron and rudder. Thus there is a possibility of many feedbackloops. The implementation of some of these can be viewed simply as synthetic modi-fication of the inherent stability derivatives. For example, p fed back to aileron modi-

Page 289: Stability Control EtkinReid

8.6 Lateral Control 281

6oo.-------------------------.

500

400

300

200

100

-100 /-

: f-200 : \\! ex 2000,rad-3000L..--.L--..J---~-....L----'--~---'-:--:-:--~:--~200

20 40 60 80 100Time,s

(a)

200.--------------------------,

lie X 2000,rad100~/ r-----------,"-"- T= Tmax

\ I \0lrt-/~=:t:====::;;::;=;o-\-__:::::-:----____j

I\V I \ /I I \

-100i I Y5 X 1000 \../

-200--\ I\ I'- L-- T = 0

-300-

-400'-- 1'-- ---J1'-- -J1 ---'o 50 100 150 200

Time,s(b)

Figure 8.18 Altitude-hold controller, (a) Height, speed, and pitch angle. (b) Elevator and throttlecontrols.

fies Lp (roll damper), r to the rudder modifies N; (yaw damper), and v to rudder mod-ifies the yaw stiffness Nu' and so on. It is a helpful and instructive preliminary to adetailed study of particular lateral control objectives to survey some of these possiblecontrol loops. We could do this analytically by examining the approximate transferfunctions given in Chap. 7. However, we prefer here to do this by way of example,using the now familiar jet transport, and using the full system model. We treat eachloop as in Fig. 8.19, as a negative feedback with a perfect sensor and a perfect actua-tor, so that the loop is characterized by the simple constant gain K. For each case wepresent a root locus plot with the gain as parameter (Fig. 8.20) (All the root loci aresymmetrical about the real axis; for some, only the upper half is shown). As is con-

Page 290: Stability Control EtkinReid

282 Chapter 8. Closed Loop Control

Figure 8.19 Representative loop.

ventional, the crosses designate the open loop roots (poles) and the circles the openloop zeros. The pair of complex roots corresponds to the Dutch Roll oscillation; thereal root near the origin is for the spiral mode; and the real root farther to the left isthat of the heavily damped roll mode.

Since the root loci always proceed from the poles to the zeroes as IKI increases,the locations of the zeros can be just as important in fixing the character of the loci asthe locations of the poles. The numbers on the loci are the values of the gain. Zerogain of course corresponds to the original open loop roots. The objective of control isto influence the dynamics, and the degree of this influence is manifested by theamount of movement the roots show for small changes in the gain. We have not in-cluded root loci for acceleration feedback, and of the remaining ten, two show verysmall effects, and are therefore not included either. These two are the aileron feed-backs: v ~ 5a and r ~ 5a• Each of the other eight is discussed individually below.

</J~ 5a It was pointed out in Chaps. 2 and 3 that airplanes have inherentaerodynamic rotational stiffness in pitch and yaw, but that there isno such stiffness for rotations about the velocity vector. This funda-mental feature of aerodynamics is responsible for the fact that air-planes have to sideslip in order to level the wings after an initial rollupset. This lack can be remedied by adding the synthetic derivative

L", = Laa d5ald</J = - KLaaWe might expect that making such a major change as adding a newaerodynamic rotational stiffness would have profound effects on theairplane's lateral dynamics. Figure 8.20a shows that this is indeedthe case. The time constants of the two nonperiodic modes are seento change very rapidly as the gain is increased, until with even asmall gain, 1K1 < I, that is, less than 10 of aileron for 10 of bank,these two modes have disappeared, to be replaced by a low fre-quency, heavily damped oscillation. The Dutch Roll remains virtu-ally unaffected by the aileron feedback for any modest gain.This root locus is shown in Fig. 8.20b. The largest effect is on theroll mode, as might be expected, where a positive gain of unity (cor-responding to a decrease in ILpl) results in a substantial reduction inthe magnitude of the large real root. This is accompanied by an in-crease in the spiral stability and a slight reduction in the Dutch Rolldamping. A negative gain, (an increase in ILpl) increases the DutchRoll damping, shortens the roll mode time constant and causes aslight reduction in the magnitude of the spiral root (the latter notvisible in the figure).Because l/J is the integral of r, the transfer functions for l/J have an sfactor in the denominator, and hence a pole at the origin. This is

Page 291: Stability Control EtkinReid

8.6 Lateral Control 283

1.2r------------------__..;;:::----,-----,

1.0Dutch roll

0.2 -1

-0.5 a

Spiral

0.8

'" 0.6Cl

'"E 0.4

a K=-o.50.0I-"*"----_l-- __ --+ - __ ----~---""I

Roll-0.2 L- __ ---1. ....L ...u... .1...-__ ---L -'- __ ----'

-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 a 0.1Real s

(a).~l)a

1.2

f- -4 1 4..•.•.~- ~5f-

t-

t-

t-4 5

K = 1 ,,- ~-

I I I I I

1.0

0.8

'" 0.6Cl

'"E 0.4

0.2

0.0

-0.2-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 a 0.1

Real s

(bl p~l)a

Figure 8.20 Root loci. (a) c/>-+ 8a• (b) P -+ 8a• (c) l/J-+ 8Q• (d) v -+ 8,. (e) p -+ 8,. if) r-+ 8,. (g)c/>-+ 8,. (h) l/J-+ 8" (i) STOL airplane; r-+ 8,.

seen in Fig. 8.20c. The expansion theorem (A.2,10) shows that thezero root of the characteristic equation leads to a constant in the so-lution for l/J. This is consistent with the fact that the reference direc-tion for l/J is arbitrary.

The feedback of l/J to aileron has little relative effect on theDutch Roll and rolling modes. Its main influence is seen on the spi-ral and zero roots, which are quite sensitive to this feedback. Fornegative gain (stick left for yaw to the right) these two modesrapidly combine into an oscillatory mode that goes unstable by thetime the gain is -0.5 (0.50 aileron for 10 yaw). For all positivegains, there is an unstable divergence.This feedback (Fig. 8.20d) represents rudder angle proportional tosideslip, with positive gain corresponding to an increase in Nv• Note

Page 292: Stability Control EtkinReid

284 Chapter 8. Closed Loop Control

0.2 r-------------,--------..",...------,

0.1

(/)

Ol

§ OJ------ •...---~~~~--- .•.-------~K= 0.05

-0.1

-o.2.'"::----:-':-------I..----I...- ...L__ .......:s.. ---.J-0.5 -0.03 -0.D1 -0.01

Reals

(e) '1'-) lia

0.03 0.05

••1.8 r------------------~---.,------____,

0.2-0.005Ot-*'"--+--o--- ...•.--------------H!*->o-o--....,

-0.2 '-- __ ---'- --'- J-- __ ---L ----'- -'-- __ ----'

-0.6

1.4

(/)

E0.6

,

0.005 K=-o.005 -0.001 0.001

-0.5 -0.4 -0.3 -0.2 -0.1 o 0.1Real s

(d) 1) -) Ii,

Figure 8.20 (Continued)

that a gain of 0.001 for v corresponds to 5.1f3 = -0.774. The princi-pal effect is to increase the frequency of the Dutch Roll while simul-taneously decreasing the spiral stability, which rapidly goes unstableas the gain is increased. The reverse is true for negative gain. Theroll mode remains essentially unaffected.Roll rate fed back to the rudder has a large effect on all three modes.For positive gain (right rudder for roll to the right) the damping ofthe Dutch Roll is increased quite dramatically-it is quadrupled fora gain of about 0.20 rudder/deg/s of roll rate. This is counterintuitive(see Exercise 8.10). At the same time, the damping of the roll modeis very much diminished, and that of the spiral mode is increased.With further increase in gain the two nonperiodic modes combine toform an oscillation, which can go unstable at a gain of about 0.4.The large effects shown in Fig. 8.20! for the yaw damper case arewhat would be expected. As an aid in assessing the damping perfor-

Page 293: Stability Control EtkinReid

1.2,..----------------------,-------------,

U>

g> 01----+-----Q--4----:t:---.,,;..------------jE

1.2

1.0

0.8

U> 0.6

'"§ 0.4

0.2

0.0

-0.2-0.6

0.8

0.4

-0.4

-0.8

8.6 Lateral Control 285

--- Ko=0.4 0.3 0.2 ~.-.-

-

I-

'-0.3

0.40.2 .. <,

'S. ./

I I I I I

-0.5 -0.4 -0.3 -0.2 -0.1 o 0.1Real 5

(e) p --+ Ii,

-1.2L..- ..J..... ----'- ----L -'-- ---'

-1.5 -1.0 -0.5 0 0.5 1.0Real 5

if) r--+ Ii,

Figure 8.20 (Continued)

mance, two lines of constant relative damping ( are shown on thisfigure. Negative gain corresponds to left rudder when yawing noseright. A very large increase in Dutch Roll damping is attained with again of -1, at which point there is a commensurate gain in the spi-ral damping. There is some loss in damping of the roll mode. Thebeneficial effects of yaw-rate feedback are clearly evident from thisfigure. The behavior for larger negative gains, beyond about - 1.4,is especially interesting. For this airplane at this flight condition, thetwo real roots combine to form a new oscillation, the damping ofwhich rapidly deteriorates with further increase in negative gain.This feature complicates the choice of gain for the yaw damper. Thepattern shown is not the only one possible. Figure 8.20i is the corre-sponding root locus for the STOL airplane of Sec. 8.9, flying at10,000 ft and 200 k. It illustrates the importance of the location ofthe zeroes of the closed loop transfer function. For the jet transportthe real zero, z is to the left of the real roll mode root p. For the

Page 294: Stability Control EtkinReid

286 Chapter 8. Closed Loop Control

1.8

1.6

1.4

1.2

"' 1.0=KOJ 0.8Ctl

§0.6

0.4

0.2-0.01

0

-0.2-0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1

Real 5

(g) cjl -t 1),

1.2-0.5

1.0

0.50.8

"' 0.6OJCtl

§ 0.4

-0.50.2 -0.1

0

-0.2-0.05 -0.04 -0.03 -0.02 -0,01 0 0,01 0.02 0.03 0.04 0.05

Real 5

(hI 'If -t 1),

0.03

0.02

0.01

~I~ -6 -6OJ 0 ""Ctl /§ P. pole Z, zero

-0.01

-0.02

-0.03-0.05 -0.04 -0.03 -0.02 -0.01 0

Real Sc2V

(i) STOL Airplane; r -t 1),

Figure 8.20 (Continued)

Page 295: Stability Control EtkinReid

8.7 Yaw Damper 287

STOL airplane the reverse is the case. The direction of the locusemanating from this root is therefore opposite in the two cases, witha consequent basic difference in the qualitative nature of the dynam-ics. For the STOL airplane the Dutch Roll root splits into a real pair,one of which then combines with the spiral root to form a new low-frequency oscillation. In viewing Fig. 8.20i it should be noted that itwas drawn for the nondimensional system model, and hence the nu-merical values for the roots and the gains are not directly compara-ble with those of Fig. 8.20f.

Feeding back bank angle to the rudder produces mixed results (Fig.8.20g). When the gain is negative, the spiral mode is rapidly drivenunstable. On the other hand if it is positive, to improve the spiral,the Dutch Roll is adversely affected.

The consequence of using heading to control the rudder is alsoequivocal. If the gain is positive (heading right induces right rudder)the null mode becomes divergent. If the gain is negative the twononperiodic modes form a new oscillation at quite small gain thatquickly becomes unstable.

8.7 Yaw DamperYaw dampers are widely used as components of stability augmentation systems(SAS); we saw the potential beneficial effects in Fig. 8.20f. At first glance the yawdamper would appear to be a very simple application of feedback control princi-ples-just use Fig. 8.20/ as a guide, select a reasonable gain, and add a model for theservo actuator/control dynamics. However, it is not really that simple. There is an-other important factor that has to be taken into account-namely that during a steadyturn, the value of r is not zero. If, in that situation, the yaw damper commands a rud-der angle because it senses an r, the angle would no doubt not be the right oneneeded for a coordinated turn. In fact during a right tum the yaw damper would al-ways produce left rudder, whereas right rudder would usually be required (see Fig.7.24). This characteristic of the yaw damper is therefore undesirable. To eliminate it,the usual method is to introduce a high-pass or "washout" filter, which has zero gainin the steady state and unity gain at high frequency. The zero steady state gain elimi-nates the feedback altogether in a steady tum. The system that results is pictured inFig. 8.21, where the meaning of the filter time constant is illustrated. For the servoactuator/rudder combination of this large airplane we assume a first order system oftime constant 0.3 sec.

The closed loop transfer function for Fig. 8.21 is readily found to be [see (8.2,1)]

G = JGrilr

rTc 1 + WJGrilr

(8.7,1)

This transfer function was used to calculate a number of transient responses to illus-trate the effects of J(s) and W(s).

As a reference starting point, Fig. 8.22 shows the open loop response (W = 0) to

Page 296: Stability Control EtkinReid

288 Chapter 8. Closed Loop Control

r, +0-:-. s,J ««-

Rudder and airframeservo actuator

W

r

J= 3.333Ks + 3.333

filter w= _s_s+a

Figure 8.21 Yaw damper.

a unit impulse of yaw rate command r.: It is evident that there is a poorly damped os-cillatory response (the Dutch Roll) that continues for about 2 min and is followed bya slow drift back to zero (the spiral mode). Fig. 8.22a shows that the control dynam-ics (i.e., J(s» has not had much effect on the response.

Figure 8.23 shows what happens when the yaw damper is turned on with thesame input as in Fig. 8.22. It is seen that the response is very well damped with eitherof the two gains shown, which span the useful range suggested by Fig. 8.20j, and thatthe spiral mode effect has also been suppressed.

It remains to choose a time constant for the washout filter. If it is too long, thewashout effect will be insufficient; if too short, it may impair the damping perfor-mance. To assist in making the choice, it is helpful to see how the parameter a =lITwo affects the lateral roots. Figure 8.24 shows the result for a gain of K = -1.6.The roots in this case consist of those shown in Fig. 8.20! plus an additional smallreal root associated with the filter. It is seen that good damping can be realized forvalues of a up to about 0.3, that is, for time constant T down to about 3 s. This resultis very dependent on the gain that is chosen. While the oscillatory modes are behav-ing as displayed, the real roots are also changing-the roll root decreasing in magni-tude from -2.31 at a = 0 to -1.95 at a = 0.32. The new small real root starts at theorigin when a = 0 and moves slowly to the left, growing to -0.00464 at a = 0.32.When the filter time constant is 5 s the small root is -0.0038, corresponding to anaperiodic mode with thalf = 182 s. It is instructive to compare the performance of theyaw damper with and without the filter for an otherwise identical case. This is donein Fig. 8.25. It is seen that the main difference between them comes from the smallreal root, which after 5 min has reduced the yaw rate to about 5% of its peak value.This slow decay is unlikely to present a problem since the airplane heading is in-evitably controlled, either by a human or automatic pilot. In either case, the residual rwould rapidly be eliminated (see Sec. 8.8).

••

Page 297: Stability Control EtkinReid

8.7 Yaw Damper 289

0.5 f----+-----+----+------t------t-------l

-10 10 20 30 40 50 60Time, s

(a) Initial response, 0-60 s

0.5

0.25en:0~jI;l 0 I~ ~.3: IV' ,co>-

-0.25

-0.5 F-----==----__+_

With servoactuator

-0.50 100 200Time, s

(b) Long-term response

300 400

Figure 8.22 Yaw rate impulse response-open loop, W == O.(a) Initial response, 0-60 sec. (b)Long-term response.

8.8 Roll Controller

This example is of another common component of an AFeS, a control loop thatmaintains the wings level when flying on autopilot, or that can be commanded to rollthe airplane into a tum and hold it there. We shall see in this particular case that theresulting tum is virtually truly banked, even though no special provision has beenmade to control sideslip.

The block diagram of the system is shown in Fig. 8.26. It incorporates the yawdamper described in the previous section and adds two additional loops. The outerloop commands <p. The <p error is converted to a roll rate command by Jp, and it is theroll rate error that is then used to drive the aileron servo actuator. If the roll rate fol-lowed the command instantaneously, without lag, the bank angle response would beexponential (i.e., cb oc <p). In reality of course this ideal behavior is not achieved be-cause of the airframe and servo dynamics.

Page 298: Stability Control EtkinReid

290 Chapter 8. Closed Loop Control

o

Gain. k

(\-2;-1

",

J ~~ .•.• ::::---------....- -V·5 10 15

Time. 5

20 25 30-0.50

Figure 8.23 Effect of yaw damper.

For this example we use the state vector approach to system modeling in order toprovide another illustration, one that differs in detail from that of Sec. 8.5.

As usual the starting point is the basic aircraft matrix equation,

t = Ax + Be (8.8,1)

in which x = [v p r </J]T and c = [5a 5r]T.The differential equations that correspond to the various control transfer func-

tions in the figure are found as follows. For the yaw damper components, we have thesame form of transfer functions as previously, that is,

s (8.8,2)

and

W(s) = --1-s+-

TwO

K,hrJr(s) = --1-

s+-Tr

(8.8,3)

(8.8,4)

From (8.8,4) we get the differential equation

8r= Kr r _ (2- + _1_) ~r __ 1_

r, r; TwO T,TwO

For Jp we use the constant Kp, and for Ja we use a first order servoactuator

Ka/Tai,= --1-

s+-Ta

The relation between Ba, P, and lPc is seen from the diagram to be

Ba = Ja(pc - p) = JaJp(lPc - lP) - JaP

(8.8,5)••

(8.8,6)

(8.8,7)

Page 299: Stability Control EtkinReid

8.8 Roll Controller 291

~= 0.5;I0.8

Gain K= -1.6 1.0

0.6 1m

0.4

0.2

-0.9 -0.8 -0.7 -0.6 -5.0 -0.4 -0.3 -0.2 -0.1Re

Figure 8.24 System poles for varying washout time constant.

o

s 0.5

~J!i~s~ 0

IGainK= -1.6

a = 0.24

\ With washout

~ II

........ '-.... -"'1>----V '\

Withoutwashout

-0.5 o 5 10 15 20Time, s

(a) Short time, 0-30 s

25 30

0.5

0.25

~'Ce

•~ 0

s~

-0.25

GainK=-1.6a = 0.24

~

Withwashout/'

50 250 300-0.5

0 100 150 200Time, s

(b) Long time

Figure 8.25 Effect of washout filter on yaw damper performance. (a) Short time, 0-30 sec. (b)Longtime.

Page 300: Stability Control EtkinReid

292 Chapter 8. Closed Loop Control

~ell.

+~ ~-. Ja-

G

lipr. = 0 ----:0-- s, ~

W

Figure 8.26 Roll control system.

When we substitute for Ja and Jp the differential equation that results is

• KaKp s; 1 KaKp8a = - -- cP - - p - - 8a + -- cPc'Ta 'Ta 'Ta 'Ta

(8.8,8)

Equations (8.8,5) and (8.8,8) are the additional equations required to augment the ba-sic system (8.8,1) to accommodate the addition of the two control angles as depen-dent variables. However, a little more manipulation is needed of (8.8,5). To put it infirst-order form, we define the new variable

(8.8,9)

and to put it in canonical form, we must eliminate i,This we do by using the thirdcomponent equation in (8.8,1). When these steps have been taken the system can beassembled into the matrix equation

z = Pz + QcPc (a)(8.8,10)

where z = [v p r cP 8a 8r y]T (b)

The matrices P and Q are:

a11 a12 a13 g b11 b12 0

a21 a22 a23 0 b21 b22 0

a31 a32 a33 0 b31 b32 0

0 1 0 0 0 0 0

P== s, _ KaKp 10 00 0

'Ta 'Ta 'Ta

0 0 0 0 0 0 1

s; s, s,0

s:(K

r 1 ) _(~ + _1 )-a31 -a32 -a33 -b31 -b ---'Tr t; r; r, t; 32 'Tr'Twa r; 'Twa

Q ==[0 0 0 0 (KaKi'Ta) 0 O]T (8.8,12)

Equation (8.8,10) was solved by numerical integration for two cases, with the resultsshown on Figs. 8.27 and 8.28. The various gains and time constants used were se-

v

p

r

,

...

..

Page 301: Stability Control EtkinReid

8.8 Roll Controller 293

0.4

-0.2

o 5 10 15Time,s

(a) •• P, and lia

30

0.05 ,...-------------------------,

~~o

-0"E .-- ••.." 0 ~-~~~-""':7.~::om-=:.a--------------.,

-0.050 15 20

Time. 5

(bl~, r,lj1,lir

Figure 8.27 Response of roll controller to initial cp of 0.262 rad (15°). (a) cp, p, and ~a' (b) {3,T, l{J,~r

5 10 25 30

lected somewhat arbitrarily, as follows:

Kp = 1.5 Ka = -1.0 K; = -1.6; 'Ta = .15 'Tr = .30 'TwO = 4.0

On the first of these figures, response to an initial bank error, we see that all the statevariables experience a reasonably well damped oscillatory decay, and that the maxi-

Page 302: Stability Control EtkinReid

294 Chapter 8. Closed Loop Control

. .

0.4 f-

0.2 f- .•/ ..-<: x 2

o t/ -.:..-------------------------------

",...-- .•_--_ .•- .•.", 5a,,

-0.2.... "I,

I "I II Iv

-0.4 f-

I I I I I0 5 10 15 20 25 30

Time, S

(a) eIl,P, Ilo

.!!?"'0"'0~~<01.;;;. •..ai

0.1r---------"'- ..-.....-....-.....--:....,.....-------------,

."..•...-"'-/

..•.......

,~~~~~---------,-:i~~--------------------------.a: Jllr /~.-4:" . """"'-l::" e:o .••.••_ .

0.05 -

I I I I Io 5 2510 15

Time, S

20 30

(bl~, r,"', s,Figure 8.28 Response of roll controller to roll command of 0.262 rad (15°). (a) cP, p, 8

0, (b) {3,r,

I/!,8,.

mum control angles required are not excessive-about 20° for the aileron and lessthan 1° for the rudder. The time taken for the motion to subside to negligible levels isequal to about two Dutch Roll periods. All the variables except r/J subside to zero,whereas r/J asymptotes to a new steady state. When level flight is reestablished, theairplane has changed its heading by about 1.8°.

The second figure shows the response to a 15° bank command. The new steadystate is approached with a damped oscillation that takes about 15 s to decay. Thesteady state is clearly a turn to the right, in which r has a constant value and r/J is in-

,

Page 303: Stability Control EtkinReid

8.9 GustAlleviation 295

creasingly linearly. All the other variables, including the two control angles, are verysmall. It is especially interesting that the sideslip angle is almost zero. Clearly thiscontroller has the capability to provide the bank angle needed for a coordinated turn.(The angle of attack and lift would of course have to be increased.)

8.9 Gust AlleviationFor the final example in this chapter, we tum to a study of the application of auto-matic controls to reduce the response of an airplane to atmospheric turbulence(Byrne, 1983). This is obviously a useful goal for many flight situations, the benefitsincluding increased passenger comfort, reduction of pilot workload, and possibly re-ductions in structural loading and fatigue, and in fuel consumption. The case reportedhere is for a STOL airplane, which is especially vulnerable to turbulence, since therelatively low operating speed makes it more responsive to turbulence, and becauseits duty cycle requires it to spend relatively more time at low altitudes where turbu-lence is more intense.

The numerical data used in the study were supplied by the de Havilland AircraftCo., and although it does not apply to any particular airplane, it is representative ofthe class.

In this situation, where random turbulence produces random forces and momentson the airplane, which in tum result in random motion, the methods of analysis wehave used in the foregoing examples, being essentially deterministic, are not applica-ble. Random processes have to be described by statistical functions. Let f(t) representsuch a random function. Two of the key statistical properties that characterize it arethe spectrum function, derived from a Fourier analysis of f(t) and the closely relatedcorrelation function (Etkin, 1972). The spectrum function or spectral density, as it isfrequently called, is denoted <l>ff(w), The area under the <l>ff(w) curve that is co~tained between the two frequencies Wi and W2 is equal to the contribution to V2

(where P is the mean-squared-value of f)3 that comes from all the frequencies in theband Wi ~ W2 that are contained in the Fourier representation of f.

A BASIC THEOREM

When a system with transfer function G(s) is subjected to an input with spectrumfunction <I>;;(w) the spectral density <l>rr(w) of the response r(t) is given by

<l>rr(w) = <I>;;(w)IG(iw)12 (8.9,1)

where G(iw) is the frequency response function defined in Sec. 7.5. There is a gener-alization of (8.9,1) available for multiple inputs (Etkin, 1972, p. 94). Figure 8.29shows the relationships expressed in (8.9,1) for a second-order system of moderatedamping.

In the case at hand the motion studied is the lateral motion, and the forces neededare Y,L, and N. The gust vector g (Fig. 8.1) that is the source of these forces has ele-ments that represent aspects of the motion of the atmosphere. It has four components

g = [vg Pg rig r2gY

3The factor icomes from the use of two-sided spectra.

Page 304: Stability Control EtkinReid

296 Chapter 8. Closed Loop Control

0.5

4r-----------------------,

1.5

3.5i(t) r(t)

G(8)

3

2.5

2f------~···············..········..••......

/.'\ \ <Pi; (W)IG(iw)l2 " I \ /'" " ." -,' , ..•.-------------------- \ \

I \

'a \.\ '.I ••••

o '-- ..l...... -.l.... __ -=_..._...._....J.•••_••_•••_•••..;;•••::.::•••:.::;•••""••••••••-J

0.01 0.1 1.0 10m (rad/s)

Figure 8.29 Response to random input.

100

vg is the y-component of the turbulent velocity, Pg is the lateral gradient of the z-com-ponent, Pg = aw/ay, and rIg = -au/ay, r2g = av/ax. Each of these inputs is capa-ble of producing aerodynamic actions on the airframe. vg acts just like v in producingforces and moments like Yvv and Lvv; Pg acts like P and produces moments like LTiJand NTiJ; and rIg and r2g also produce forces and moments as a result of their effectson the relative wind at the wing and tail. The details of this theory can be found inEtkin (1981). The incremental force and moments can be expressed in terms of thegust components by the equation

[~l~Fg

Where F is a (3X4) matrix of "gust derivatives." The basic differential equation ofthe system is then (8.8,1) with an added term to account for the turbulence, that is,

\(8.9,1)

x = Ax + Be + Tg (8.9,2)

Tg=

o'"

In order to alleviate the response of the airplane, it is necessary to be specific andchoose an output that is to be minimized, and then to find what control actions will

Page 305: Stability Control EtkinReid

8.9 Gust Alleviation 297

9

Kg

G'(s) J> x+

e c,/ J(s) G(s)

+

Kx

Figure 8.30 Gust alleviation system.

be successful in doing so. In the cited study, the output chosen to be minimized was apassenger comfort index (see below).

A block diagram of the system considered is given in Fig. 8.30. The state vectoris the set x = [v p r 4JJT and the control vector is c = [Sa 8rJT. The model includesfull state feedback via the (2X4) gain matrix Kx, control servo actuators described bythe (2X2) matrix J, and also includes the possibility of using measurements of theturbulent motion to influence the controls via the (2X4) gain matrix Kg.

We now proceed to complete the differential equation of the system. We startwith the servo actuator transfer function J(s), which is given, as in our previous ex-amples, by first-order elements:

lITa0

S + liraJ= (8.9,3)

Ur;0

S + lITr

Equation (8.9,3) corresponds to the pair of differential equations

Sa = (ea - Sa)lTas, = (e, - 8r)lTr

or C = -Pc + Pe

(8.9,4)

(8.9,5)

10

where P= Ta

1(8.9,6)

0Tr

We see from Fig. 8.30 that e == Kgg + KxX, so that (8.9,5) becomes

c = PKxX - Pc + PKgg (8.9,7)

Page 306: Stability Control EtkinReid

298 Chapter 8. Closed Loop Control

We can now combine (8.9,2) and (8.9,7) into the augmented differential equation ofthe system:

(8.9,8)

z=Az+tgThis can be written more compactly, with obvious meanings of the symbols, as

(8.9,9)

In the cited study, various control strategies were examined, differentiated primarilyby whether or not gust "feedforward" was included (i.e., Kg *" 0). When only statefeedback was employed, (Kg = 0) linear optimal control theory was used to ascertainthe optimum values of the gains in Kx• To this end, a function has to be chosen to beminimized. The choice made was a passenger comfort index made up of a linearcombination of sideways seat acceleration along with angular accelerations p and f.The seat acceleration depends on how far the seat is from the CO, so an average wasused for this quantity. The optimum that resulted entailed the feedback of each of thefour state variables to each of the two controls, a very complicated control system!However, it was found that there was very little difference in performance between

0.30

0.20

0.25

Fixed/contro!s

"boS.e•....oo 0.15

0.10

0.05

0.01

"

Page 307: Stability Control EtkinReid

8.9 GustAlleviation 299

this optimum and a simple yaw damper. The result is shown on Fig. 8.31, in the formof the spectral density of sideways acceleration of the reannost seat. This form ofplot, f$(f) vs. log f, is commonly used. The area under any portion of this curve isalso equal to the mean-square contribution of that frequency band, just as with $(f)vs. f. Results are shown for three cases-the basic airframe with fixed controls, aconventional autopilot, and the selected yaw damper. Very substantial reduction of re-sponse to turbulence has clearly been achieved with a relatively simple control strat-egy.

An alternative to conventional linear optimal control theory was found to be bet-ter for the case when gust measurement is assumed to be possible. It stems from atheorem of Rynaski et al. (1979). It is seen from (8.9,2) that if one could make Be +Tg = 0 then one would have completely canceled the gust input with control action,and the airplane would fly as if it were in still air! This equality presupposes that thecontrol is given by

c = -B-1Tg

that is, that B has an inverse. This would require B to be a square matrix [i.e., to be(4X4)], which in turn would require that the airplane have two more independent

0.10

With gustalleviation

0.30

0.25

Fixed/controls

0.20

"be

~.•...g 0.15

0.05

0.00 L_+_-==:::::;;;;;iiiiI!!~~~=::::::r:;_~0.01 0.1 1.0

Frequency, hz

Figure 8.32 Lateral acceleration spectra. Rearrnost seat, with yaw damper and gust feedforward.

Page 308: Stability Control EtkinReid

300 Chapter 8. Closed Loop Control

(and sufficiently powerful) controls than it actually has. Although this is not beyondthe realm of imagination, it was not a feasible option in the present study. However,there is available the "generalized inverse," which provides in a certain sense the bestapproximation to the desired control law. The generalized inverse of B is the inverseof the (2X2) matrix BTJI. This leads to the control law

c = _(BTJI)-lBTTg

(The second BT is needed to yield a (2X 1) matrix on the right-hand side). This lawstill requires, however, that all four components of g be sensed in order to compute c.Sensing all components of g is not impossible, indeed it may not even be impractical.However, a good result can be obtained with a subset of g consisting only of vg andT2g, both of which can be measured with an aerodynamic yawmeter, a sideslip vane orother form of sensor. The end result of combining gust sensing in this way with theyaw damper, with the gust sensor placed an optimum distance forward of the CG, isshown in Fig. 8.32. It is clear that this control strategy has been successful in achiev-ing a very large reduction in seat acceleration.

8.10 Exercises8.1 Assume that an aerodynamic derivative L", has been added to the lateral force system.

What changes does this entail in the lateral characteristic equation? What implica-tions do these changes have for lateral dynamics?

8.2 (a) What is the steady state ()that results from a steady t15e = 5° for the jet transportof Sec. 8.3?

(b) For the closed-loop response to a unit step input in Sec. 8.3, with J = l0., derivean expression for the steady-state error ess as a function of l0.. (Hint: start with(8.3,1)).

(c) Calculate the value of l0. needed to keep ess < 0.1° for ()c = 5°.(d) For the value of k2 found in (c) what is the elevator angle at t = 0+ when ()c is a

step input of 50? Comment on the practicality of using l0. alone to reduce ess•

8.3 (a) With respect to Fig. 8.5, write out the transfer function for the elevator angle re-sponse to ()c input.

(b) Calculate the steady-state response for the case of Fig. 8.7c.

8.4 (a) Derive the expressions for the transfer functions Gu6p and Gy8p given in (8.4,2).(b) Derive the expressions for the closed loop transfer functions given in (8.4,3).

8.5 The system of Fig. 8.8 is to be represented by the block diagram of Fig. 8.1 with x =[u w q OYand c = [5e 5pY.Write out the matrices D, E, and H. What are the di-mensions of J?

8.6 Given the 2X2 algebraic system

Ai=Bc

Page 309: Stability Control EtkinReid

8.11 Additional Symbols Introduced in Chapter 8 301

for which

Xi Nis)cj D(s)

where D(s) is det A. Prove that

det N = det A det B

that is, that

Relate this result to the elimination of fin (8.4,6).

8.7 Modify the system model of the altitude-hold autopilot (Fig. 8.17) to include a rateterm in the block indicated by the constant k. Show explicitly what changes result in(8.5,13).

8.8 Modify the analysis of Sec. 8.5 that leads to (8.5,13) to include climbing or glidingflight, that is, (Jo =1=0.

8.9 (a) Prove that in the yaw damper with washout, the steady-state yaw rate for a step inrc is independent of the washout time constant and is given by

rss = r,j(O)GraJO)

(b) Prove that if the washout filter is in the forward path, instead of the feedbackpath, then

regardless of the washout time constant.

8.10 A positive gain K in the p ~ 5r loop (Figs. 8.19 and 8.20e) implies right rudder re-sponse when the right wing dips down. On the face of it this would seem to be desta-bilizing. Explain, using the modal diagram Fig. 6.15 and Table 6.9, why the dampingof the Dutch Roll mode is actually increased by this feedback. (Hint: r produced by ayaw damper is normally used to increase Dutch Roll damping.)

8.11 Add an outer loop to the system of Fig. 8.26 to control the heading angle ljJ. Draw anew block diagram and write out the augmented system differential equation. (Hint:design the loop to command a bank angle proportional to heading error.)

8.12 Write out the full system of equations corresponding to (8.9,8).

8.13 (This is a miniresearch project). The relative locations of the pole P and the zero Z onFig. 8.20f are relevant to the design of a yaw damper in that they determine the char-acteristic structure of the root locus. Discover, by any means, what changes to the jettransport's aerodynamics would move its zero to the right of the pole. Two sugges-tions: (1) Systematically perturb each of the stability derivatives and note the sensi-tivity of the locations of P and Z to each, and (2) use approximate transfer functionsto get analytical approximations for P and Z. When you have found some changes tothe stability derivatives that would produce the desired result, discuss the designchanges that would be needed to achieve the altered derivatives.

Page 310: Stability Control EtkinReid

302 Chapter 8. Closed Loop Control

8.11 Additional Symbols Introduced in Chapter 8

c control vectorD(s) denominator of transfer functionDv aD/aVe error vectorF matrix of gust derivatives .•g gust vectorG' transfer function g :::}x

G closed-loop transfer functionH feedback matrix

J transfer function e :::}c

K gain constantN(s) numerator of transfer function

Pg gust gradient awlay

rIg gust gradient -au/ayrZg gust gradient av/axr reference vectorT gust input matrix

Tv aTiavV* speed at which Tv = Dv

ug, Vg, Wg components of gust velocity

y output vector ••z feedback vectorcf>( ltJ) spectrum function

'T time constant

Page 311: Stability Control EtkinReid

APPENDIX A

Analytical Tools

A.I LinearAlgebraIn this book no formal distinction is made between vectors and matrices, the formerbeing simply column matrices, as is common in treatments of linear algebra. In par-ticular the familiar vectors of mechanics, such as force and velocity, are simply three-component column matrices. We use boldface letters for both matrices and vectors,for example, A = [au] and v = [vJ The corresponding lowercase letter defines themagnitude (or norm) of the vector. The transpose and inverse are denoted as usual bysuperscripts, for example, AT and A-t. When appropriate to the context, a subscriptis used to denote the frame of reference for a physical vector, for example, VE =[u V wf denotes a vector whose components in frame FE are (u, v, w). The three-component vectors of physics have the following properties:

Scalar product

(AI,l)

c is a scalar, with magnitude ab cos (J, where (J is the angle between a and b

Vectorproduct

(A 1,2)

where[

0 -a3 a2]

i = a3 0 -at-a2 at 0

(A1,3)

c is a vector perpendicular to the plane of a and b, with direction following theright-hand rule for the sequence a, b, c and has the magnitude ab sin (J, where()is the angle (<180°) between a and b

Unit vectorsThe basis unit vectors are i,j,k such that

[

1 0[i j k] = 0 1

o 0(A 1,4)

303

Page 312: Stability Control EtkinReid

304 Appendix A. Analytical Tools

Where 5ij is the Kronecker delta.Square matrices have the following properties:

Minor Determinant, Cofactor

The minor determinant mij of a matrix [aij] is the determinant of the reduced ma-trix that remains after the ith row andjth column of [aij] have been deleted.The cofactor is cij = mij( -IY+j

Adjoint

The adjoint of a matrix is the transpose of the matrix of cofactors,

adj [aij] = [cijt (AI,5)

Inverse

Provided that det A '* 0 the inverse is given by

A-I = adj AdetA

(AI,6)

A.2 The Laplace TransformLet x(t) be a known function of t for values of t > O.Then the Laplace transform ofx(t) is defined by the integral relation

xes) = .;E[x(t)] = f' e-stx(t) dto

(A2,1)

The integral is convergent only for certain functions x(t) and for certain values of s.The Laplace transform is defined only when the integral converges. This restriction isweak and excludes few cases of interest to engineers. It should be noted that the orig-inal function x(t) is converted into a new function of the transform variable s by thetransformation. The two notations for the transform shown on the left-hand side of(A2,1) will be used interchangeably. The transforms of some functions that com-monly occur in problems of linear systems are listed in Table AI.

TRANSFORMS OF DERIVATIVES

.;E [dx J = L"" e-st dx dtdt 0 dt

= J.oo e-st dx = xe-stJoo + s foo xe:" dtt=O t=O Jo

When xe -st ~ 0 as t ~ 00 (only this case is considered), then

.;E [: J = -x(O) + sx(s) (A2,2)

,"

Page 313: Stability Control EtkinReid

A.2 The Laplace Transform 305

TableA.lLaplace Transforms

x(t) i(S)

1 8(t) 1

2 8(t - T) «:"

13 1 or l(t) -

ae-sT

4 l(t - T) -a

5 f(t - T)1(t - T) e -sT 5£[f(t)]

16 t

a2

r:' 17 -

(n - I)! an

18 e" --

a-a

a9 sin at

a2 + a2

a10 cos at

~ +a2

111 te" (a - a)2

r:' 112 eat

(n - I)! (s - a)n

e" sin btb

13 (s - a)2 + b2

e" cos bta-a

14(a - a)2 + b2

a15 sinh at

~- a2

s16 cosh at

a2 - a2

17 e" sinh btb

(a - a)2 - b2

18 e" cosh bta-a

(s - a)2 - b2

19 i(t) ai(s) - x(0)

20 e"'x(t) i(a - a)

Page 314: Stability Control EtkinReid

306 Appendix A. Analytical Tools

where x(O) is the value of x(t) when t = 0.1 The process may be repeated to find thehigher derivatives by replacing x(t) in (A2,2) by x(t), and so on. The result is

[dnx] dn-1x dn-2x:£ - = - -- (0) - s -- (0) - ... - sn-lx(O) + snx(s) (A 2 3)dt" dtn-1 dr'"? . ,

TRANSFORM OF AN INTEGRAL

Let the integral be

y = f x(t) dt

and let it be required to find yes). By differentiating with respect to t, we get

dydi = x(t)

thus

xes) = :£ [ ~] = sy(s) - yeO)

and

1 1yes) = - xes) + - yeO)

s s(A2,4)

METHODS FOR THE INVERSE TRANSFORMATION

The Use of Tables of TransformsExtensive tables of transforms (like Table AI) have been published that are use-

ful in carrying out the inverse process. When the transform involved can be found inthe tables, the function x(t) is obtained directly.

The Method of Partial FractionsIn some cases it is convenient to expand the transform xes) in partial fractions, so

that the elements are all simple ones like those in Table AI. The function x(t) canthen be obtained simply from the table. This procedure is illustrated with an example.Let the second-order system of Sec. 7.3 be initially quiescent, that is, x(O) = 0, andx(O) = 0, and let it be acted upon by a constant unit force applied at time t = O.Thenf(t) = 1, and !(s) = lis (see Table A.I). Then (see (704,1»

1xes) = 2 2 (A2,5)

s(s + 2lwns + Wn)

'To avoid ambiguity when dealing with step functions, t = 0 should always be interpreted ast = 0+,

Page 315: Stability Control EtkinReid

A.2 The Laplace Transform 307

Let us assume that the system is aperiodic; that is, that C > 1. Then the roots of thecharacteristic equation are real and equal to

AI,Z = n ± w' (A.2,6)

where

n = -Cwnw' = wn(Cz - 1)112

The denominator of (A.2,5) can be written in factored form so that

1xes) = ------

s(s - AI)(S - Az)

Now let (A.2,7) be expanded in partial fractions,

ABCxes) = - + + ---

S (s - AI) (s - Az)

By the usual method of equating (A.2,7) and (A.2,8), we find

1A=-~

AIAz1

B=-----\1(-\1 - Az)

1C=----

AzCAz - AI)

(A.2,7)

(A.2,8)

Therefore

lIAIA2 lIAI(AI - Az) lIAzCAz - AI)xes) = -- + + ---'----s S - Al S - Az

By comparing these three terms with items 3 and 8 in Table A.l, we may write downthe solution immediately as

(A.2,9)

Heaviside Expansion TheoremWhen the transform is a ratio of two polynomials in s, the method of partial frac-

tions can be generalized. Let

N(s)xes) = D(s)

Page 316: Stability Control EtkinReid

308 Appendix A. Analytical Tools

where N(s) and D(s) are polynomials and the degree of D(s) is higher than that ofN(s). Let the roots of D(s) = 0 be a., so that

D(s) = (s - aI)(s - a2) ... (s - an)

Then the inverse of the transform is

~ {(S - ar)N(s) }x(t) = L eartr=I D(s) s=ar

The effect of the factor (s - ar) in the numerator is to cancel out the same factor ofthe denominator. The substitution s = a, is then made in the reduced expression?

In applying this theorem to (A.2,7), we have the three roots al = 0, a2 = AI'a3 = A2, and N(s) = 1.With these roots, (A.2,9) follows immediately from (A.2,10).

(A.2,1O)

The Inversion Theorem

The function x(t) can be found formally from its transform i(s) by the applicationof the inversion theorem Jaeger (1949) and Carslaw and Jaeger (1947). It is given bythe line integral

1 iY+iWx(t) = -2 . liII!e esti(s) ds

7rl w-' y-iw(A.2,1l)

where 1'is a real number greater than the real part of all values of s for which i(s) di-verges. That is, s = l' is a straight line on the s plane lying parallel to the imaginaryaxis, and to the right of all the poles of i(s). This theorem can be used, employing themethods of contour integrals in the complex plane, to evaluate the inverse of thetransform.

Extreme Value TheoremsEquation (A.2,2) may be rewritten as

-x(O) + si(s) = (' e-sti(t) dt

= lim iT e-stx(t) dtT-.oo 0

We now take the limit s ~ 0 while T is held constant, that is,

-x(O) + lim si(s) = lim iTlim e-sti(t) dt.•.....•.0 T-+oo 0 .•.....•.0

I= lim iT i(t) dt = lim [x(T) - x(O)]

T-+oo 0 T-+oo

Hence lim si(s) = lim x(T).•.....•.0 T-+oo

(A.2,12)

This result, known as the final value theorem, provides a ready means for determin-ing the asymptotic value of x(t) for large times from the value of its Laplace trans-form.

2Por the case of repeated roots, see Jaeger (1949).

Page 317: Stability Control EtkinReid

A.3 The Convolution Integral 309

In a similar way, by taking the limit s ~ 00 at constant T, the integral vanishes forall finite x(t) and we get the initial value theorem.

lim si(s) = x(O) (A.2,13)~'"

A.3 The Convolution IntegralThe response of any linear system to any arbitrary input f(t) can be obtained from in-tegrals of the two basic response functions h(t) and A(t). h(t) is the response to theunit impulse 8(t), and A(t) is the response to the unit step l(t). The system is assumedto be initially quiescent. If not, the transient associated with nonzero initial condi-tions must be added to the following integrals. The response to f(t) is then given byDuhamel's integral, or the convolution integral:

x(t) = r h(t - T)f( T)dr7=0

x(t) = r A(t - T)j( T)dr7=0

(a)

(A.3,1)

(f(O) = 0) (b)

When f(O) is not zero, then there must be added to (A.3,Ib) a term to allow forthe initial step in f(t); i.e.,

x(t) = f(O)A(t) + f:o

A(t - T)j( T)dr (A.3,2)

The physical significance of these integrals is brought out by considering them as thelimits of the following sums

x(t) = Lh(t - T)f( T) aT

x(t) = A(t)f(O) + LA(t - T)j(T) aT(a) (A.33)(b) ,

Typical terms of the summations are illustrated in Figs. A.I and A.2. The summationforms are quite convenient for computation, especially when the interval aT is keptconstant.

'I

t

Figure A.I Duhamel's integral, impulsive form, ~ = h(t - T)f( T) aT = response at time t toimpulse at time T.

Page 318: Stability Control EtkinReid

310 Appendix A. A1Ullytical Tools

/:.{=f'I'TI/:.'T

1,--:==-.11"-t - l/:.';Staircase" representation of {It)

/:.'T

:1Figure A.2 Duhamel's integral, indicial form. I1f = step input applied at time T, Ax =A(t - T)l1f = response at time t to step input 11f.

A.4 Coordinate TransformationsTRANSFORMATION OF A VECTOR

Let v be a vector with the components

[

Vbl]and Vb = vln in Fb

Vb3

The component of val in the direction of Xb/ is Val cos (Oil) where Oil denotes the an-gle between 0bXb/ and 0axal (see Fig. A.3). Thus by adding the three components ofvaj in the direction of xb/ we get

i = 1 ... 3 (A.4,1)

where

lij = cos (Oij)

are the nine direction cosines. (A.4,1) is evidently the matrix product

(A.4,2)

(a)where (A.4,3)

(b)

and constitutes the required transformation formula. Its inverse readily reverses thetransformation to give

Page 319: Stability Control EtkinReid

A.4 Coordinate Transformations 311

Figure A.3 Component of vector.

where (A,4,4)

When a vector is successively transformed through several frames of reference, forexample, Fa' Fb, Fe ... then

andVb = Lbava

v, = LebVb = Leb(LbaVa)

Since also Ve = Lea Va' then it follows that

and similarly for additional transformations.The sequence of subscripts in the preceding expression should be noted, as it

provides a convenient mnemonic for remembering these relations.

PROPERTIES OF THE L MATRIX

Since va and Vb are physically the same vector v, the magnitude of va must be thesame as that of Vb' that is, v2 is an invariant of the transformation. From (A,4,3) thisrequires

(A,4,5)

It follows from the last equality of (A,4,5) that

(A,4,6)

Equation (AA,6) is known as the orthogonality condition on Lba• From (A,4,6) it fol-lows that

ILbal2 = 1

and hence that ILbal is never zero and the inverse of Lba always exists. In view of(AA,6) we have, of course, that

(A,4,7)

Page 320: Stability Control EtkinReid

312 Appendix A. Analytical Tools

that is the inverse and the transpose are the same. Equation (AA,6) together with(A.4,3b) yields a set of conditions on the direction cosines,

3

L IkJkj = 5ijk=l

(AA,8)

It follows from (A.4,8) that the columns of Lba are vectors that form an orthogonalset (hence the name "orthogonal matrix") and that they are of unit length.

Since (A.4,8) is a set of six relations among the nine lij' then only three of themare independent. These three are an alternative to the three independent Euler anglesfor specifying the orientation of one frame relative to another.

THE L MATRIX IN TERMS OF ROTATION ANGLES

The transformations associated with single rotations about the three coordinate axesare now given. In each case Fa represents the initial frame, Fb the frame after rota-tion, and the notation for L identifies the axis and the angle of the rotation (see Fig.A.4). Thus in each case

(A.4,9)

By inspection of the angles in Fig. A.4, the following matrices are readily verified.

LI(Xl) = [~ .: x, .: z, ]o -sin x, cos x,

(A.4,1O)

\xta %oa \SIIl %01

\ \\ \\ \\ \\ \ ••.~SII.\ ••.~S6:I \ ......\ ...... \ ...... x•Xl \

......\ ...... ......•..... •.. %0••.. %cit0 0

(0) (6)

\%6,. %as\\\\ ...••..~

\ .\ ..•."••.....,- Xa"':;;---"""--%01

(e)

Figure A.4 The three basic rotations, (a) Aboutxa1, (b) Aboutxa2, (c) Aboutxa3,

Page 321: Stability Control EtkinReid

A.4 Coordinate Transformations 313

The transformation matrix for any sequence of rotations can be constructed readilyfrom the above basic formulas. For the case of Euler angles, which rotate frame FEinto FB as defined in Sec. 404, the matrix corresponds to the sequence (X3, X2, Xl) =(l/J, (), cP), giving

(A.4,1l)

[The sequence of angles in (A.4, 11) is opposite that of the rotations, since each trans-formation matrix premultiplies the vector arrived at in the previous step.] The resultof multiplying the three matrices is

cos ()cos l/J cos ()sin l/J -sin ()

sin cP sin ()cos l/J sin cP sin ()sin l/J sin cP cos ()LBE= - cos cP sin l/J + cos cP cos l/J (A.4,12)

cos cP sin ()cos l/J cos cP sin ()sin l/J cos cP cos ()+ sin cP sin l/J - sin cP cos l/J

TRANSFORMATION OF THE DERIVATIVE OF A VECTOR

Consider a vector v that is being observed simultaneously from two frames Fa and Fb

that have relative rotation-say Fb rotates with angular velocity Cd relative to Fa'which we may regard as fixed. From (A.4,3)

The derivatives of Va and Vb are of course

and (Ao4,13)

where val = (d/dt)(va), and so forth. It is important to note that va and Vb are notsimply two sets of components of the same vector, but are actually two different vec-tors.

Now because Fb rotates relative to Fa' the direction cosines Iij are changing withtime, and the derivative of (A.4,3) is

(A.4,14)

or alternatively

the second terms representing the effect of the rotation.Since L must be independent of v, the matrix Lab can readily be identified by

considering the case when Vb is constant (see Fig. A.5.). For then, from the funda-mental definitions of derivative and cross product, the derivative of vas seen from Fais readily shown to be

dv-=wXvdt

(Ao4,15)

Page 322: Stability Control EtkinReid

314 Appendix A. Analytical Tools

Figure A.5 Rotating vector of constant magnitude.

The matrix equivalent to (A.4,15) is

(A.4,16)

where

The corresponding result from (A.4,14) is

va = t:»,It follows from equating (A.4,16) and (AA,17) that

(AA,17)

or

(A.4,18)

for all Vb. Whence

andLab = 6)aLab

6)a=L~ba

Finally if the above argument is repeated with Fb considered fixed, and Fa having an-gular velocity -~, we clearly arrive at the reciprocal result

(A.4,19)

From (A.4,18) and (AA,19), recalling that 6) is skew-symmetric so that 6)T = -6),the reader can readily derive the result

(AA,20)

Page 323: Stability Control EtkinReid

A.S Computation of Eigenvalues and Eigenvectors 315

From (A.4,14), (A.4,18), and (A.4,19) we have the alternative relations

Vb = Lbava - iiJbVb

Va = LabVb + iiJava(A.4,21)

with two additional permutations made possible by (A.4,20). A particular form weshall finally want for application is that which uses the components of Va transformedinto Fb, viz.

(A.4,22)

TRANSFORMATION OF A MATRIX

Equation (A.4,20) is an example of the transformation of a matrix, the elements ofwhich are dependent on the frame of reference. Generally the matrix of interest A oc-curs in an equation of the form

v=Au (A.4,23)

where the elements of the (physical) vectors u and v and of the matrix A are all de-pendent on the reference frame. We write (A.4,23) for each of the two frames Fa andFb, that is,

va = AauaVb = Abub

(a)(A.4,24)

(b)

and transform the second to

Premultiplying by Lab we get

Va = La~~baUa

By comparison with (A.4,24a) we get the general result

Aa = La~~ba

(A.4,25)

(A.4,26)

A.5 Computation of Eigenvalues and EigenvectorsSome software packages provide for the calculation of eigenvalues and eigenvectorsof matrices directly. The software used for many of the computations in this book isthe Student Version of Program CC,3 which does not do this. However, these impor-tant system properties can readily be obtained from it, as shown in the following.

Program CC is oriented to the calculation of transfer functions and presents themin various forms; one is the pole-zero form. Any transfer function of the system, forexample, that from elevator angle to pitch rate, when displayed in this form, willshow the eigenvalues in the denominator. That is how we obtained the eigenvaluespresented in Chap. 6.

3Available from Systems Technology, Inc., 13766 South Hawthorne Blvd., Hawthorne, CA, 90250-7083 U.S.

Page 324: Stability Control EtkinReid

316 Appendix A. Analytical Tools

For the eigenvectors, we turn to the expansion theorem (A.2,1O). Consider a casewhere the input to the system is Sc = S(t), Dirac's delta function. The response of theith component of the state vector to this input in the mode corresponding to eigen-value A is

X.(t) = [ (s - A)Nj(s) ] e"• D(s) s=A.

The ratio of this component to Xl for the same input S(t) is

xj(t) Nj(A)--=--xl(t) NI(A)

This ratio gives the ith component of the eigenvector for the mode associated with A.Any component can be chosen for reference instead of Xl' as illustrated in Figs. 6.3and 6.15.

(A.5,1)

(A.5,2)

A.6 Velocityand Acceleration in an ArbitrarilyMoving Frame

Since in many applications, we want to express the position, inertial velocity, and in-ertial acceleration of a particle in components parallel to the axes of moving frames,we need general theorems that allow for arbitrary motion of the origin, and arbitraryangular velocity of the frame. These theorems are presented below.

Let F~ Oxyz) be any moving frame with origin at 0 and with angular velocity (JJ

relative to Fl' Let r = ro + r' be the position vector of a point P of FM (see Fig. A.6).Let the velocity and acceleration of P relative to FI be v and a. Then in FI

VI = tl (A.6,1)

We want expressions for the velocity and acceleration of P in terms of the compo-nents of r' in FM' Expanding the first of (A.6, 1)

VI = tal + t;= Val + t;

(A.6,2)

"Figure A.6 Moving coordinate system.

Page 325: Stability Control EtkinReid

A.6 Velocity and Acceleration in an Arbitrarily Moving Frame 317

where Vo is the velocity of 0 relative to Fj• The velocity components in FM are givenby

VM = LMjvj = LMj(vOj + t;) = YOM+ LMjt;

From the rule for transforming derivatives (A.4,22)

(A.6,3)

whence(A.6,4)

The first term of (A.6,4) is the velocity of 0 relative to Fj, the second is the velocityof P as measured by an observer fixed in FM, and the last is the "transport velocity,"that is the velocity relative to F, of the point of FM that is momentarily coincidentwith P. The total velocity of P relative to F, is the sum of these three components.Following traditional practice in flight dynamics, we denote

(A.6,5)

(When necessary, subscripts are added to the components to identify particular mov-ing frames.)

The scalar expansion of (A.6,4) is then

Vx = vox + X + qz - ry

vy = VOy + Y + rx - pz (A.6,6)

Vz = VOl + Z + py - qx

These expressions then give the components, parallel to the moving coordinate axes,of the velocity of P relative to the inertial frame.

On differentiating Vj and using (A.6,4) we find the components of inertial accel-eration parallel to the FM axes to be

8M = LMjYj = YM + iiJMVM= YOM+ r~ + 6)Mr~ + iiJMt~ + iiJMvOM+ iiJMt~ + iiJMiiJMr~

= 80M + r~ + 6)Mr~ + 2iiJMt~ + iiJMiiJMr~ (A.6,7)

where 80M = YOM+ iiJMvOM= LMjyO/ is the acceleration of 0 relative to Fj.The total inertial acceleration of P is seen to be composed of the following parts:

80M: the acceleration of the origin of the moving framer~: the acceleration of P as measured by an observer fixed in the mov-

ing frameGJMr~: the "tangential" acceleration owing to rotational acceleration of the

frameFM

2iiJMt~: the Coriolis accelerationwM6JMr~: the centripetal acceleration

Three of the five terms vanish when the frame FM has no rotation, and only r~ re-mains if it is inertial. Note that the Coriolis acceleration is perpendicular to wM and

Page 326: Stability Control EtkinReid

318 Appendix A. Analytical Tools

t;", and the centripetal acceleration is directed along the perpendicular from P to (x).

The scalar expansion of (A.6,7) gives the required inertial acceleration componentsof Pas

ax = aox + oX + 2qz - 2ry - x(tf + "z) + y(pq - t) + z(pr + q)ay = aoy + Y + 2ri - 2pz + x(pq + t) - y(p2 + "z) + z(qr - p) (A.6,8)

az = aoz + Z + 2py - 2qi + x(pr - q) + y(qr + p) - Z(p2 + q2)

Page 327: Stability Control EtkinReid

APPENDIX B

Data for EstimatingAerodynamic Derivatives

This appendix contains a limited amount of data on stability and control derivatives.It is not intended to be used for design. That requires much more detail than couldpossibly be provided here. It is intended to display some representative orders ofmagnitude and trends, and to provide numerical data that teachers and students canuse for exercises. All the data pertain to subsonic flight of rigid airplanes. Much ofthe information comes from either the USAF Datcom (USAF, 1978) or from the datasheets of the Royal Aeronautical Society of Great Britain (now out of print), which isalso the source for some of the Datcom data. We have taken some liberties in extract-ing and presenting this information, but have not changed any essential content. Forinformation about derivatives at transonic and supersonic speeds and for geometriesdifferent from those covered in the following, the reader is referred to the USAF Dat-com. When estimating derivatives, reference should also be made to Tables 5.1 and5.2.

B.I Lift-Curve Slope, CLa

B.2 Control Effectiveness, CL8

B.3 Control Hinge Moments

B.4 TabEffectiveness, b3

aEB.5 Downwash, aa

B.6 Effect of Bodies on Neutral Point and Cmo

B.7 Propeller and Slipstream Effects

s.s Wing Pitching Derivative, c.;B.9 Wing Sideslip Derivatives C1p, c.,B.IO Wing Rolling Derivatives C1p' c;B.II Wing YawingDerivatives C1r, c;

B.12 Changes in Inertias and Stability Derivatives withChange of Body Axes

319

Page 328: Stability Control EtkinReid

320 Appendix B. Data for Estimating Aerodynamic Derivatives

NOTATION

AC

CLa

<(Ct)theory

KMR

/3K

NOTES

B.l Lift-Curve Slope, CLa

aspect ratiochordlift-curve slope of wing alonetwo-dimensional (airfoil) lift-curve slopetheoretical value of Ciaan empirical factorMach numberReynolds number, VcptJ.t

Prandtl-Glauert compressibility factor, VI - M2

/3Ctl27Tsweepback angle of midchord line

• The source of the data for airfoils and wings is USAF datcom. It applies torigid straight-tapered wings at subsonic speeds and small angle of attack.

• The section lift-curve slope is given by

1.05Cia = T K( CI)theory (B.1,1)

where K is given in Fig. B.1,la and (Ct)theory in Fig. B.1,lb. Y90 and Y99 are the air-foil thicknesses, in percent of chord, at 90% and 99% of the chord back from theleading edge, as illustrated, and the trailing edge angle is defined in terms of thesethicknesses by

(B.1,2)

• The lift-curve slope Cia of the wing alone is given in Fig. B.1,2. The insetequation is seen to approach the theoretically correct limits of 7TA/2 as A ~ 0and 27Tas {A~ 00, K~ 1, A~ 0, /3~ l}.

• Figure B.1,3 gives some theoretical values of the body effect on CL., forunswept wings in mid-wing combination with an infinite circular cylinderbody. For values of A < 1, the theory also applies to delta wings with pointedtips.

In Fig. B.1,3a the wing angle of attack is the same as that of the fuselage; that is,e = O. In this case the lift of the wing-body combination increases to a maximumvalue, then decreases with increasing body diameter. Where there is a wing setting,i.e., e =1=0, and as = 0 (Fig. B.1,3b), the lift of the combination decreases with in-creasing o:

Page 329: Stability Control EtkinReid

B.2 Control Effectiveness, CLs 321

Ygg

~--y~~2

1

~

!-0.09C "I1.0.------r---,.-----r---,.-----r---r--...,..--r--...,..----,

.9 "'=-+---+--p""-d---+-=.......q;;;:::--+----+---f------j

K

.8 1---+---+----1'-----+-~....l:---t--+--+--....,._.;;;:_-j

Note: Interpolate by plottingK vs 10glO R 4---+---+--"',.......:---I---/-----t

.7oL--...L.-----l..--l-----=--':-:----l...----J.=---.l...---:-:----L.":::::O"'~0.08 .12 .16

tan + ~'TE

(a)

7.4

7.2

7.0c:-o"Q)

~ ~ 6.8CDtl a.

~6.6

6.4

./

V•....

./

./ V./ V

./'" VV

0.04 0.08 .12 .16Wing thickness ratio, tic

(b)

Figure B.I,1 Two-dimensional lift-curve slope.

.20

B.2 Control Effectiveness, CLa

SECTION DATA

Figure B.2,la presents theoretical values of the two-dimensional control derivativeCis for simple flaps in incompressible flow. These values can be corrected by the em-

Page 330: Stability Control EtkinReid

322 Appendix B. Data/or Estimating Aerodynamic Derivatives

.6

r--...r--... cl I I I I I I

La 2n'\ "A= --, 2 +j2f32 (1 + tan2 :cn )+ 4 _

" r 13-,<,

i'o...••..••..........

"""'- ---r---- -I- -

1.6

1.4

1.2

1.0

'"I -rJ ~ 1" .8

.4

.2

00 2 3 4 5 6 7 8 9 10 11 12 13 14 15 161

~ [132+tan2 A"n]2

Figure B.1,2 Subsonic wing lift-curve slope.

pirical data of Fig. B.2,lb for the strong effect of nonideal Iift-curve slope of themain surface to which the control is attached.

SURFACE DATAThe derivative CLa for a finite lifting surface with a part span control flap is obtainedfrom the section derivative by

••

(CLa)CLl; = Cia C

lc, K}K2

where CLa and Cia are as defined in B.I, Cia is the corrected value from Fig. B.2,lband K1 and K2 are the factors given in Figs. B.2,2 and B.2,3. In these figures the para-meter (a6)C/ is the rate of change of zero-lift angle with flap deflection, given by theinset graph, and Aw and Aare, respectively, the aspect ratio and taper ratio of the mainsurface.

B.3 Control Hinge Moments

NOTATION

'T trailing-edge angle defined by the tangents £0 the upper andlower surfaces at the trailing edgetheoretical rate of change of hinge-moment coefficient withangle of attack for incompressible inviscid two-dimensionalflowactual rate of change of hinge-moment coefficient with angleof attack for incompressible two-dimensional flow

Page 331: Stability Control EtkinReid

B.3 Control Hinge Moments 323

0.9

0.8

0 0 0.7. •w 2:: 0.6to to.•.•...•.•..tS" tS"

0.6e e0.4

0.3

0.2

0.1

Body diameter to wing span, U = dlb(aJ

R_'''wl''''(_? c=]~;;;-infinite circular _ _ - - : ~cylinder mid-wingconfiguration

1.0 -=::=:;::~-II-rT---,rTI---'0.9

0.8

0.7

0 0 0.6• •III •• ••~ ~ 0.6.•.•..~tS"

e e0.4

0.3

0.2

0.1

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Body diameter to wing span, U = dlb(bl

Figure B.l,3 Body effect on lift-curve slope expressed as a ratio of lift of wing-body combinationto lift of wing alone. (From "Lift and Lift Distribution of Wings in Combination with SlenderBodies of Revolution," by H. J. Luckert, Can. Aero. J., December 1955.)

Page 332: Stability Control EtkinReid

324 Appendix B. Datafor Estimating Aerodynamic Derivatives

6.0 r--'T""""-"'T""-~-"""--'T--""'-"T"""--r--..,...- ..•.•

(Cl) 4.0o theory

(per radian)

6.6

6.0

'I

1('-~r-_O.150.12

0.10-t->r--->0.08~-~---1,

0.060.04

0.02+---t----+-~f____I

tic· 0

4.6t--..,....---,-----,--.....---..,....~

3.5t----+---+---+--

(a)

3.01--+--_+_--"2I

oo~_...L-- ....•..- ...•._.....L----I-----:L-_..L-_ ...•...._...I-_...J0.1 0.2 0.3 0.4 0.5

2.6t--+----:ff----t---t----j--t--+---1----j----1

2.0/---tt:....---j---+--t----j--t--+---1----j----1

~~ --r----:::::~-- ~ -~ :::::~ ---- -~ l..--- ,...---- -- -V.--- --~ f...-- ~ f..--

~ f...-- -.--- -- ~ ~ f..--- -- - -f.--V-- ~ -- L---~--.---I...----~ -~

CIO

(C'O)theOry 0.6

1.0

0.9

0.8

0.7

0.5

0.4

1.0 0.2 0.3

C,1.00= -20.98 (Clu) theory0.960.940.920.900.880.860.840.820.800.780.760.740.720.70

••

(h)

0.4 0.5erie

Figure B.2,! Control effectiveness for two-dimensional incompressible flow. (From RoyalAeronautical Society Data Sheet Controls 01.01.03.)

.'

theoretical rate of change of hinge-moment coefficient withcontrol deflection for incompressible inviscid two-dimen-sional flowactual rate of change of hinge-moment coefficient with con-trol deflection for incompressible two-dimensional flow

Page 333: Stability Control EtkinReid

B.3 Control Hinge Moments 325

2.0 -1.0

-.8

1.8 -.6(lX5)C

Z-.4

1.6 -.2

K,.2 .4 .6 .8 1.0

eric

1.4

AwFigure B.2,2 Flap-chord factor.

rates of change of control hinge-moment coefficients with in-cidence and control-surface deflection, respectively, in two-dimensional flow for control surfaces with sealed gap andnose balance

induced angle of attack correction to (b1)o and (b2)o, respec-tively, where F1 is the value of (a/B) [CZ/CZa] when cf = C

stream-line curvature correction to (b1)o and (b2)o, respec-tively, where F2 is the value of A(b2) when cf = C

factor to F2 and A(b2) allowing for nose balanceratio of control-surface area forward of hinge line to control-surface area behind hinge line

F3

Balance

NOTES

Figures B.3,! and B.3,2The curves of Fig. B.3,1 were derived for a standard series of airfoils with plain

controls for which tan (~)'T = tic (referred to by an asterisk). To correct for airfoils

Page 334: Stability Control EtkinReid

326 Appendix B. Data/or Estimating Aerodynamic Derivatives

..

1.0

1.0

.8

~~

......:A.

~ ~0/j ~

-. 5

~ ~ "1.0.I- ~

#1$

~

VI---

~v

.6

.4

.2

o o .2 .4 .6 .8 1.01\ = y/J!....

2

Figure B.2,3 Span factor for inboard flaps.

with tan (i)T different from tic, values of (bl)~T>(C1);'COry and C;a are calculated forthe given tic ratio; then (b1)o is calculated from

(B.3,l)

Values of (C1);'cory and C;a may be obtained as in Appendix B.l.The curves apply for values of angle of attack and control deflection for which

there is no flow separation over the airfoil; for these conditions (b1)o can be estimatedto within ±O.05. The data refer to sealed gaps but may be used if the gap is notgreater than O.OO2c.

The above discussion also applies to the data given in Fig. B.3,2 for (b2)o' Thesubscript 1 in Eq. B.3,l becomes a subscript 2, a becomes l>, and values of (C1J;'COryand C;8 may be obtained from Appendix B.2.

Page 335: Stability Control EtkinReid

B.3 Control Hinge Moments 327

0.9

0.8

0.7

0.6e.!!

~ 0.6~c.~ 0.40~..,.

T 0.3

0.1

00 0.1 0.2 0.3 0.4ere

1.0 1.000.980.96

0.8 0.940.920.90

... 0.6 0.880 0.86s,

~ 0.84s i 0.82..... 0.40 .,!-;; 0.80~§ •-I-o ~ 0.78

0.2 0.760.74

0 0.72

0.70

o 0.1 0.2 0.3 0.4ereFigure B.3,] Rate of change of hinge-moment coefficient with angle of attack for a plain controlin incompressible two-dimensional flow. (From Royal Aeronautical Society Data Sheet Controls,04.01.01.)

Figure B.3,3

The effect of nose balance on (b1)o and (b2)o can be estimated from the curvesgiven on this figure. The data were obtained from wind-tunnel tests on airfoils withcontrol-chord/airfoil-chord ratio of 0.3. Relatively small changes in nose and trailing-edge shape, and airflow over the control surface, may have a large effect on hingemoments for balanced control surfaces, so that estimates of nose-balance effect willbe fairly inaccurate. If the control-surface gap is unsealed, the hinge-moment coeffi-cients of plain and nosebalanced controls will generally become more positive.

Page 336: Stability Control EtkinReid

328 Appendix B. Data/or Estimating Aerodynamic Derivatives

1.10r--.---T'"'"-~-"""-""'-""""-"""'---k- .~~ 0v : I I---+--....j...,~~ 0.04

"'I 0.080.08 ~

0.96t--.----r---r-....,----:~~_b..~~~~ 0.10 ••0.12

~ 0.16•• 0.90 r--t--t-::'7'"F--h~~~H~~+.l8-8

~ 0.86~--r---+""7"9-"""'7'~~~I£-++-~-I----l.()

T

1.00

1.06

0.76r--r---b"e....--il,e--1~~I--+---I----l

0.80t--t--'f£--7"f-rh'£"'h~+--+---l

0.70 t--t--f---btC--+--+---+----I-~

o 0.1 0.3 0.4

1.0

--r--..-r-- -r--.t--r--- r----.."-'"""'-

'" r-,•.....

""" <,<

1.000.950.90

0.86

0.80

0.9

100 0.8.0

to

~ 0.7o•.g

0.6

0.75 COlo

0.70 (c,;;r) theory

0.650.5

0.60o 0.1 0.2 0.3 OA

ctcFigure B.3,2 Rate of change of hinge-moment coefficient with control deflection for a plain flapin two-dimensional flow. (From Royal Aeronautical Society Data Sheet Controls 04.01.02.)

Figure B.3,4Two-dimensional hinge-moment coefficients for control surfaces with nose bal-

ance can be corrected for finite aspect ratio of the main surface using the factorsgiven in the curves and the following equations:

bI = (bIMI - FI) + F2F3Cla

b2 = (b2)o - (a/5)(bI)o + a(b~F3CI8

(B.3,2)(B.3,3)

Page 337: Stability Control EtkinReid

B.3 Control Hinge Moments 329

1.0

0.8

J c 0.6

i.2

~ oS 0.4oS

0.2

00

List of svmbols

o NACA 0009 }• NACA 0015 Nose-shapeo NACA 66009 round+ NACA 0009 > EIII •x NACA 0015 pne/I NACA 0009 Sherp

Elliptic

Round0.1 0.2 0.3 0.4 xO.5 0.6

Belanceratio = Ilcb/cfl2 - It/2cr'2)1/2

0.7

1.0

0.8

J c0.6:ia.

~.2 oSoS 0.4

0.2

00 1.0 0.2 0.3 0.4 0.5 0.6 0.7

Balanceratio

Figure B.3,3 Effect of nose balance on two-dimensional plain-control hinge-momentcoefficients. (From Royal Aeronautical Society Data Sheet 04.01.03.)

For plain control surfaces the above equations are used with F3 = 1. (b1)o and (b2)ocan be obtained from Fig. B.3,1 and B.3,2, respectively, for plain controls. For nose-balanced controls, the two-dimensional coefficients (b1)o and (b2)o must include theeffect of nose-balance. Values of Cia can be obtained from Sec. B.l, and those for Ciafrom Fig. B.2,1.

Lifting-surface theory was applied to unswept wings with elliptic spanwise liftdistribution to derive the factors. Full-span control surfaces were assumed togetherwith constant ratios of c/c and constant values of (b1)o and (b2)o across the span. Thefactors apply to wings with taper ratios of 2 to 3 if c Jc, (b1)o and (b2)o do not vary bymore than ± 10% from their average values.

Page 338: Stability Control EtkinReid

330 Appendix B. Datafor Estimating Aerodynamic Derivatives

0.6r---,---r-...,..-...,..--r--~-""""'""",,-"""--

~ 0.3r--4~~~~:::-+----;r---+-+-+-+---Id'Ic3'°

cr/oO 0.2/--t--f--+-......::::

0.1t--t---lr---1----4--+--4----l---+.---+.---1

O~2--3:--•.4-.....J.5---J6---'7L--a.l.--gJ.....-1.L0--I.11-.....J12

A~Cia

'\

'6e!r::ci. 0.02r--~~k:---+--+---+--+--+--4-~----I':9.<I

3 4 5 6 7 a 9 10 11 12

A~Cia

Figure B.3,4 Fmite-aspect-ratio corrections for two-dimensional plain and nose-balanced controlhinge-moment coefficients (CIa per radian). (From Royal Aeronautical Society Data Sheet Controls04.01.05.)

B.4 TabEffectiveness, b3

The data and method that follows is taken from the USAF Datcom. It provides esti-mates of b

3[see (2.5,1)] for two-dimensional subsonic attached flow over airfoils

with a control surface and tab. Corrections for part-span tabs can be made by multi-plying the result for two dimensions by the ratio of the control surface area spannedby the tab to the total control surface area (both areas being measured aft of the hingeline).

•••

Page 339: Stability Control EtkinReid

B.4 Tab Effectiveness, b3 331

where

is the change in control section hinge-moment coefficient due to tabdeflection, measured at constant values of lift and flap deflection.This value is obtained from Fig. BA,l.

is the change in control section hinge-moment coefficient due to liftvariation, measured at constant values of tab and flap deflection.This value is obtained from Fig. BA,2.

is the section lift-curve slope of the primary panel (wing, horizontaltail, etc.) at constant values of tab and flap deflection. This valuecan be obtained from (B.I,I).

is the rate of change of angle of attack due to a change in tab deflec-tion in the linear range at constant values of lift and flap deflection.This value can be obtained from Fig. BA,3.

-.016r---,---.,----,---.,----,---.,----,---,-----.------,

-r-.....

-.oos/---+---t---+---t---+---+---+--+---+-----1

-.004f------1------1~ c_ ~_

f----+--/- -Experimental(NACA0009airfoil-round nose, sealed gaps)

I I I I I I Io0'----'-----.2'----'----A'----'----.L6

--L---.LS

--L----J1.0

Cflc

Figure B.4,1 Effect of tab deflection on control-surface section hinge moments.

Page 340: Stability Control EtkinReid

332 Appendix B. /Jat%r Estimating Aerodynamic Derivatives

o -. -.I'"

<,

"I\..

""<, -,

••

-.0

-.0

-.1

••-.1

-.1o .2 .4 .6 .8 1.0crlc

Figure B.4,2 Effect of section lift coefficient on flap section hinge moments.

deB.5 Downwash, aU

The method and data that follow are taken from the USAF Datcom. The average low-speed downwash gradient at the horizontal tail is given by

OE _ 1/2 1.19oa - 4.44 [KAKAKH (cos Ac/4) ] (B.5,1)

oo .2 .4 .6 .8 -1.0

-.21\\,

~

\I\.t-,

r-,<,

<,<,---

••...-.4

~----.~I~-- -.6

-.8

-1.0

Figure B.4,3 Rate of change of angle of attack due to a change in tab deflection.

Page 341: Stability Control EtkinReid

aEB.S Downwash, aa 333

.5 I I I IKA=J....- -'-

A , +A1.7

\\\,

r-,r-... r--....r-- --

.4

.3

.2

.1

2 4 6 8 10A

Figure B.S,1 Wing aspect-ratio factor.

where KA, K),o and KH are wing-aspect-ratio, wing-taper-ratio, and horizontal-tail-lo-cation factors obtained from Figs. B.5,l, B.5,2, and B.5,3, respectively. Acl4 is thesweepback angle of the wing t chord line.

At higher subsonic speeds the effect of compressibility is approximated by

(B.5,2)

1.4r-, I I I

KA= 'O-3A

'", 7-,~ <,

<,r-,<,r-, -,r-,

.2 .4 .6 .8 1.0

1.3

KA1.2

1.1

1.-

Figure B.S,2 Wing taper-ratio factor.

Page 342: Stability Control EtkinReid

334 Appendix B. Data/or Estimating Aerodynamic Derivatives

1.4 ,-----r--,...--..-----,---.---~--,....__-____._--_r_--

l-lh:1f----+---+~,----+--+---+- KH = -- ~---I-----I-----<~ 2~H

.41----+--+----1----I----+----+---I----If----J=:s::::;~

Wing

Tail maclH._~tt=hH

~Rootchord

ma~~

.2~--+-

12~HI

O'--_----JL-_---'-__ .......L__ ....L__ ..L- __ .L- __ L-_---'- __ .......L _o .2 .4 .6 .8 1.0

Figure B.5,3 Horizontal-tail-location factor.

where

( ~: )lOW speed

(CL.)low speed and (CL.)M are the wing lift-curve slopes at the appropriate Machnumbers, obtained by using the straight-tapered-wingmethod of Sec. B.I

is obtained using (B.S,I)

..

Page 343: Stability Control EtkinReid

NOTATION

e

w

e'

B.6 Effectof Bodies on Neutral Point and Cmo 335

B.6 Effect of Bodies on Neutral Point and Cmo

local wing chord at center line of fuselage or nacelle

mean aerodynamic chord

maximum width of fuselage or nacelle

gross wing area

shift of neutral point due to fuselage or nacelle as a fraction of c,positive aft

area of planform of body

area of planform of body, forward of 0.25c

root chord of wing without fillets

increment to Cmo due to a body at zero lift

reflex angle of fillet, i.e., angle between wing root chord and lowersurface of fillet for upswept fillets, or the upper surface for down-swept fillets, positive as indicated in Fig. B.8,2

fillet lift-increment ratio, i.e., C1/C1a, considering the fillet to be aflap of chord 1f

NOTES

Figure B.6,1: 411"The data for estimating !i.hn presented in this graph were derived from wind-tun-

nel tests. The forward shift in neutral point is mainly dependent on the length andwidth of the body forward of the wing. The values of !i.hn given by the curves are ac-curate to within ±O.Olc, and are about 5% higher for low-wing, and the same amountlower for high-wing configurations. The data are inapplicable if the wing is clear ofthe body. Separate values should be computed for fuselages and nacelles, and the re-sults added to obtain the total neutral-point shift.

Figure B.6,2: (CmJBThe curves given in this figure apply to stream-line bodies of circular or near cir-

cular cross section with midwing configurations. For high- or low-wing configura-tions a positive or negative !i.(Cmo)B = 0.004 is added, respectively, to the value de-rived from the curves. The curves apply only for angles of attack up to about 15° forstream-line bodies where the pitching moment of the body varies linearly with angleof attack.

In the wind-tunnel tests from which the data were derived, the wings had straighttrailing edges at the wing-body junction. Fillets have a large effect on Cmo' however,especially if ()is large. The following equation may be used to estimate the fillet ef-fect if 0.12 < lJe < 0.5 and 0.03 < SJb < 0.075:

Page 344: Stability Control EtkinReid

336 Appendix B. Data for Estimating Aerodynamic Derivatives

0.9 r----,.--,r-r--r-----------,

0.81-----+-+-1-+-1

0.71----+-+-+ .....•

0.6t----t---l..--\l--\-----,

N~ 0.5t----+--+---,I-\-~,-jI~~'1O.41-----t---+-~,........3~..,.__-_._----I

0.3t---+---\--+----1Ir-+-~'r_

0.21----+--~---+-~-t__-~o;::__---1

0.1J----+---+-~-+---t__-~o;::__---1

O~-~~-~--~--.L...-----L--.....o 0.1 0.2 0.3 0.4 0.5 0.6

••

ell

Figure B.6,1 Effect of a fuselage or nacelle on neutral-point position. (From Royal AeronauticalSociety Data Sheet Aircraft 08.01.01.)

The value obtained from the curves, the fillet effect, and the effect due to wing posi-tion are added to determine (Cmo)B'

B.7 Propeller and Slipstream EffectsPROPELLER NORMAL FORCEThe following method of estimating the propeller normal force is due to Ribner(1944). The normal force is expressed in terms of the derivative aCN/oap (see Sec.3.4), which is given by

acN/aap = tc-;The factor f is the same for all propellers, and is given in Fig. B.7,1 as a function ofT; = T/pV2tP. The value of CY~o varies with thepropeller and its operating condition.The values for a particular propeller family are given in Fig. B.7,2. Extrapolation to

Page 345: Stability Control EtkinReid

0.7 Propeller and Slipstream Effects 337-t-i~no-lift line

t--E~-?:rZero pitching·momentline of body alone

Mean aerodynamicquarter-chcrd position

-0.30 r----.,----....,....----,.---r-----,

-0.25 1----+---+-------1:: =--+---=-"0.40

0.35

0.30

0.25

0.20

Nsl~0.15

0.10

0.05

-0.20

1 •• 1 ...•••

"'~

~I -0.15

o aS!.I.-

-0.10

-o·05I=t:::::~;;;;::;;;::;_r-1~9

Figure 0.6,208.01.07.)

oO'---- ....•..---.L...-----'---- .•....-----'0.2 0.4 0.6 0.8 1.0

SBF 'BF

SB 'B

Effect of a fuselage on Cmo' (From Royal Aeronautical Society Data Sheet Aircraft

other propellers can be made by means of Fig. B.7,3, on the basis of the "side-force-factor," SFF. This is a geometrical propeller parameter, given approximately by

SFF = 525[(blD)o.3 + (blD)o.6] + 270(blD)o.9where (bID) is the ratio of blade width to propeller diameter, and the subscript is therelative radius at which this ratio is measured.

Also given in Ribner (1944) are some curves which are useful for estimating theupwash or downwash at the propeller plane. These are reproduced in Fig. B.7,4.

Page 346: Stability Control EtkinReid

338 Appendix B. Datafor Estimating Aerodynamic Derivatives

0.8Te

Figure B.7,l Variation of f with Te. (From NACA Wartime Rept. L-25, 1944, by H. S. Ribner.)

2.0

1.8

1.6

1.4

1.2

1.0

0.8

0.6-0.4

-I--"••.•...••....

-'1./V

""V

1/1/,

••

o 0.4 1.2 1.6 2.0

LIFT DUE TO SLIPSTREAM

The method of Smelt and Davies (1937)* can be used to estimate the added wing liftdue to the slipstream. It is given by

DIesc, = S s(ACLo - 0.0008)

where

D1 = diameter of slipstream at the wing C.P.= D[(1 + a)/(1 + S)]I/2

C = wing chord on center line of slipstream

0.6 1 I ID'.} -~ rotation 6 blad

I./'~V Single es

rotation _

L•••..••V V IJI' Single rotation- I--

JI'VV I

'/ V ".,...- """"" 4b1r

"' .-.V /'" VV -,.,- 3 bl~des

/ V•....

~ ...-.-V V """"" ---- 2 blades-/ ---V ~~ ~

~ ••

0.5

0.4

o~~ 0.3

\oJ

0.2

0.1

00 10 20 30 40 50 60 70 80(J at 0.75 R, degrees

Figure B.7,2 Variation of CroWwith blade angle. (From NACA Wartime Rept: L-25, 1944, by H.S. Ribner.)

"See also Ribner and Ellis (1972).

.•

Page 347: Stability Control EtkinReid

B.7 PropeUerand Slipstream Effects 339

1.80•.•••---T----.-----r----r-----,----r-----,

1.401-----+-----1-----1---~~:~::"---~=_~3-b-lad-l ••-, l-in-g~le-r-ot-+at""':"ion-----.,4blades.linglerotation6blad88.linglerotation

1.201-----+----+-- ......••

1.00I----+---¥-----f---+-----1'----+------l

0.801---- '-----1-----1-------1-----/-----1------.,

0.60

0.4040 18080 80 100 120 140 160Sld.forcefactor,SFF

Figure B.7,3 Ratio of normal force derivatives. (From NACA Wartime Rept. L-25, 1944, by H. S.Ribner.)

, I~I

A-12~9JA ~ ~6

~ ~~-""

~ t'l!'].

A-12

9"~~ -./-iioo"'" ~6

2.4

2.0

~.t 1.2I

0.8

0.4

o -1.2 -0.8 -0.4 o 0.4 0.8 1.2 2.0DiltlncebehindrootqUlrter-chordpoint.rootchords

Figure B.7,4 Value of 1 - dE/dO!on longitudinal axis of elliptic wing for aspect ratios 6, 9, and12. (From NACA Wartime Rept. L-25, 1944, by H. S. Ribner.)

Page 348: Stability Control EtkinReid

340 Appendix B. Datafor Estimating Aerodynamic Derivatives

2.0 -1/--

/.<1.0

o o 2 4 6Aspect ratio of wing portion in slipstream

Figure B.7,5 Empirical factor A. (From Smelt and Davies, 1937.)

S = wing areas = a + ax/(D2/4 + ~)1/2

D = propeller diametera = -i + i(l + 8Tc/'TT')1/2

x = distance of wing c.P. behind propeller

CLo = lift coefficient at section on slipstream center line, in absence of the slip-stream

ao = two-dimensional lift-curve slope of wing section

8 = angle of downwash of slipstream at wing C.P. calculated from the equa-tion

1100.8 = O.OI6x/D + 1180°.8

where

0o = at/J/(l + a)

t/J = angle between propellor axis and direction of motion.

A is an empirical constant given in Fig. B.7,5.

B.8 Wing Pitching Derivative Cmq

The method of USAF Datcom for estimating this derivative for a rigid wing in sub-sonic flow is as follows. The low-speed value (M "'" 0.2) of Cmq is given by

{

A [~ (hnw - h) + 2(hnw - h)2]

-0.7C1a cos Acl4 A 2 A+ cos cl4I ( A3 tarr' ACI4) I}+- +-

24 A+6cosAcl4 8

(B.8,1)

Page 349: Stability Control EtkinReid

B.9 Wing Sideslip Derivatives C'fl' c., 341

where

Cia is the wing section lift curve slope from Sec. B.l (per rad).Acl4 is the sweepback angle of the wing ichord line.

For higher subsonic speeds the derivative is obtained by applying an approximatecompressibility correction.

(B.8,2)

where A is aspect ratio, and

(B.8,3)

B.9 WingSideslip Derivatives c; c.,The methods that follow are simplified versions of those given in USAF Datcom.They apply to rigid straight-tapered wings in subsonic flow.

The derivative CI,,:

For A ~ 1.0:

_ [(<) (<; )1 (Clf!) tJ.Clf!Clf!- CL C KMA + C + r r KMr + ()tan Acl4 () A (per deg)L ACI2 L tan c14

(B.9,1)

ForA < 1.0:

c; = CL [- 5;.3 ~ ~] - r (~)(per deg) (B.9,2)

where

is the wing-sweep contribution obtained from Fig. B.9,1.

is the compressibility correction to the sweep contribution, obtainedfrom Fig. B.9,2.

is the aspect-ratio contribution, including taper-ratio effects, ob-tained from Fig. B.9,3.

Clf!rr

is the dihedral effect for uniform geometric dihedral, obtained fromFig. B.9,4.is the dihedral angle in degrees.is the compressibility correction factor to the uniform-geometric-di-hedral effect, obtained from Fig. B.9,5.

Page 350: Stability Control EtkinReid

342 Appendix B. Data/or Estimating Aerodynamic Derivatives

-20.002

oAc/2 (deg)

20 40 60 80

o ~I

~A.=1

~ ~~

~~ r-, A1

~ ~

~24

~ 68

OJe.g -.002

~ -.004~

---.< -.006Gca.I~---- -.008

-.010

-20.002

oAc/2 (deg)

20 40 60 80

o ~ ~A.1=.5

"""""'-=

~ ~ ...•..•.

~ s---, A

~

1

2468

-.002

-.004

---.~ca.1""l -.006G Q

------ -.008

-.010

-20.002

oAc/2 (deg)

20 40 60 80

o ~I

~=o........••

"~~ ~ e-, A

~~

11.5

'\ 236-8

-.002

l;;.a -.004

~--:--, -.006G~---- -.008

-.010

Figure B.9,1 Wing sweep contribution to elf!'

is the wing-twist correction factor, obtained from Fig. B.9,6.

o is the wing-twist between the root and tip stations, negative forwashout (see Fig. B.9,6).is the sweepback angle of the midchord line.is the sweepback angle of the i chord line.

Page 351: Stability Control EtkinReid

B.9 Wing Sideslip Derivatives C'p' c., 343

2.2 I I I IA

= 1

1

/=COS hel2

/ 81/ 1/I

/ /// /6

'/' ./ ", 5

~ ~ V 4-•.•......" ~ - 2.....f,-- _3-

2.0

1.8

1.4

1.2

.4 .6MCOSAe/2

Figure B.9,2 Compressibility correction factor to sweep contribution to wing Clp'

.2 .8 1.0

The derivative Cnp:

Cnp) __ 1_ [_1_ _ tan Acl4 (cos A _ A _ A2

c: low speed - 57.3 41TA 1TA(A+ 4 cos Acf4) cf4 2 8 cos Acf4

sin A )]+ 6(hnw

- h) A cf4

(per deg) (B.9,3)

For subcritical speeds, the low-speed derivative can be modified by the Prandtl-Glauert rule to yield approximate corrections for the first-order three-dimensional ef-fects of compressibility:

a 2Aspect ratio, A

4 6 8

a "...-- ,...--/"" V-~..-- ~ -

/-:VV

A. II/a

VI.5 II1

-.008

-.012

Figure B.9,3 Aspect ratio contributions to wing Clp'

Page 352: Stability Control EtkinReid

344 Appendix B. Data/or Estimating Aerodynamic Derivatives

CI~

r(per deg2)

-.0001

CI~

r(per deg2)

-.0001

CI~

r(per deg2)

-.0001

-.0003 I A,,/2(a) 1..= 1

0

~." ±40°

./ ---~

......-±60°-~ .---

I~ ~I---""""

VV

-.0002

o o 2 4 6Aspect ratio. A

8 10

-.0003 I(b) 1..= 0.5 A,,/2

e:::: 0

V ±40°

~c::::- ~

--' ±600

/.~V

t>V

1/

-.0002

o o 2 4 6Aspect ratio. A

8 10

-.0003 I(e) A.= 0

A,,/2

~0

t.--: t:::::: ±40°

~ --- ±60°

~ ~ L--~

s- ."p1/

-.0002

4 6Aspect ratio. A

Figure B.9,4 Effect of uniform geometric dihedral on wing C1/3'

2 8 10

(Cn/3) _ ( A + 4 cos Acl4 )( A2B2 + 4AB cos Acl4 - 8 cos

2Acl4)( Cn/3)

Ci M - AB + 4 cos Acl4 A2 + 4A COS Acl4 - 8 cos" Acl4 ci low speed

(B.9,4)

where B is given by (B.8,3).

Page 353: Stability Control EtkinReid

B.lO Wing RoUing Derivatives C'p' c.; 345

1.8

I I_A_ =10COS he/2 I

/8//II 6

/

/' /./ ~

,/VV4-~ b:::=--f-- I..--- 2

1.6

1.2

1.0 0 .2 .4 .6 .8 1.0

MCOSAe/2

Figure B.9,S Compressibility correction to dihedral effect on wing C,fJ•

B.lO Wing Rolling Derivatives c., Cnp

The following methods are simplified versions of those given in USAF Datcom. Theyapply to rigid straight-tapered wings in subsonic flow, in the linear range of CL vs. a.

The derivative C1p:

(RIO,I)

-.00005

I ~.6-1.~/ .4

:::::V //

# •.... II/

/ /v

¥ ---- ---./...-

Root-section/zero-lift line

-

~- --~ --e (deg) --~ •

__ Tip-section -zero-lift line

I J J I I J

2 4 6 8 10Aspect ratio, A

Effect of wing twist on wing C,fJ•

12 14

-.00004

'ba>

"C -.00003

~~

:fi-<'':::J ~ -.00002

t::-a>

-.00001

Figure B.9,6

Page 354: Stability Control EtkinReid

346 Appendix B. Data/or Estimating Aerodynamic Derivatives

where

is the roll-damping parameter at zero lift, obtained from Fig.B.lO,! as a function of A{3 and f3AIK.The parameter K is the ratio of the two-dimensional lift-curve slopeat the appropriate Mach number to 27Tlf3; that is, (CI)M/(27Tlf3).The two-dimensional lift-curve slope is obtained from Sec. B.l.For wings with airfoil sections varying in a reasonably linear man-ner with span, the average value of the lift-curve slopes of the rootand tip sections is adequate.The parameter A{3 is the compressible sweep parameter given as

(tan AC/4)

A{3 = tan-I f3 ' where f3 = VI - M2•

and Acl4 is the sweepback angle of the wing ichord line.

is the dihedral-effect parameter given by

(Cl )r [Z ( Z)2 ]p = 1 - 2 -b sin r + 3 bl2 sin2 r

(Cl)r=o 12(B.lO,2)

where

r is the geometric dihedral angle, positive for the wing tip above the plane ofthe root chord.

(a) A.= 0.25 1MK10

'l-:::: •.. 9 -h:'-- 8 -~ 7~ ~l..-- 6

t-...I-- 5 r--r::::: ~""'"

4.5 F::::~~43.5 -::::::3 ~ I--2.5 --~~-2

~1.5 --r--.1

~~ 0 ~ ~ 00 00Ap (deg)

Figure B.IO,la Roll damping C1p' part 1.

Page 355: Stability Control EtkinReid

oII

rJ-:;:--1I;,) lo!en----

B.lO Wing RoUing Derivatives Clp' c., 347

(b) A.=0.50-.6

-.1

I3A""K10.-- 9 --....<,8

~ ~~ 7 -...- 6 r-.... .:::::~..- -5 <,~ ~- 4.5 --

4 --....:~ ~3.5 -- 1=3 -~ ~

2.5 -r-2 - Nr---r-

1.5 r--.

-.5

-.4

-.3

-.2

o-20 0 20 40 60 80

A~ (deg)

Figure B.lO,lb Roll damping C1p' part 2.

(e) A.= 1.0

10I--" r--..- 9 r-.. .::::::a- <,

7 .....•..•...-s'""" -r-. ~6 <,~-r-....

~5 ""- ...•.• '"~4.5 roo- r---.....

4 .::::::~ ~ ~3.5

3 -r-....~

~2.5

~-

2 -1.5 --~--....

1

o~O 0 ~ ~ ~ M

A~(deg)

Figure B.lO,lc Roll damping C1p' part 3.

Page 356: Stability Control EtkinReid

348 Appendix B. Datafor Estimating Aerodynamic Derivatives

z is the vertical distance between the CG and the wing root chord, positive forthe CG above the root chord.

b is the wing span.

(aCt)drag is the increment in the roll-damping derivative due to drag, given by

a (Ct)CDL c2 - ~ C( Ct)drag = Ci L 8 Do (B. 10,3)

where

is the drag-due-to-lift roll-damping parameter obtained from Fig.B.1O,2 as a function of A and Acl4•

is the wing lift coefficient below the stall.is the profile or total zero-lift drag coefficient.

The derivative Cnp:

c; = -Ctp tan ex - [-Ctp tan ex - (~: )CL=O CL] + (a~np) ()M

(B. IDA)

where

is the roll-damping derivative at the appropriate Mach number esti-mated above

ex is the angle of attack.CL is the lift coefficient.

( Cnp ) is the slope of the yawing moment due to rolling at zero lift given byCL CL=O

M

(deglJ70

60 \1\ r\\ -,

50 \ <,r-,"-

1\ -, -.............. r---40 r-, r--.1

~r-, -..r'-30 r--20 to- t-

10 ~ ....•. -t---0

2 4 6 8 10A

Figure B.I0,2 Drag-due-to-lift roll-damping parameter.

Page 357: Stability Control EtkinReid

B.lO Wing Rolling Derivatives C1p' Cnp 349

-.0010

!1Cnp

9(per deg)

o

J INote: t--

~9 in degrees

\\'\ l\\'l\\\ \\ l'"

1\ \ ,\'"\ \ r, r-,1\ \ \ r-,r"\ I'\.

\ \ \ I'"1\ 1\ -, <, A.

1"-1.0

\ \ \~

<,r-,I\. I\.

f\ \ \ r-,\ 1\ 1\ ~I\.

"\ , r-,f\ 1\-, \ \ .6I\. '\<, \ ~<,r-, -, -,

'-- Root-sectionIzero-lift line <, .2 r."'-.'---~ -~-6 deg -~

..•...

KTip-section'-- zero-lift line

I I I I I I2 4 6 8

Aspect ratio, A10 12

-.0008

-.0006

-.0004

-.0002

.0002

.0004

.0006

.0008

.0010 o

Figure B.IO,3 Effect of wing twist on wing rolling derivative Cnp'

Page 358: Stability Control EtkinReid

350 Appendix B. Data/or Estimating Aerodynamie Derivatives

where

B is given by (B.8,3).ACI4 is the sweepback angle of the t chord line.

( Cnp ) is the slope of the low-speed yawing moment due to rolling at zeroCL Ctt~g lift given by

(tan Acl4 tan2 Acl4)

(Cnp) = _ ~ A + 6(A + cos Acl4) (hnw - h) A + 12CL CL=O 6

M=O A + 4 cos Acl4(B.1O,6)

t::..Cnp()

()

is the effect of linear wing twist obtained from Fig. B.IO,3.

is the wing twist between the root and tip stations in degrees, nega-tive for washout (see Fig. B.1O,3).

B.II WingYawingDerivatives c.. c;The following methods are simplified versions of those given in USAF Datcom. Theyapply to rigid straight-tapered wings in subsonic flow at low values of CL-

The derivative c;(B.ll,l)

where

( Clr) is the slope of the rolling moment due to yawing at zero lift given by

CL CL=OM

A(I - B2) AB + 2 cos Acl4 tan2 Acl4I+ -------- + --------''-

(Clr) __ 2_B...:.(A_B_+_2_co_s_A...,;c::...:14::....)__ A_B_+_4_co_s_A...,;c....:/4__ 8__ ( Clr)

CL CL =0 = A + 2 cos Acl4 tarr' Acl4 CL CL=OM I + --- M=O

A + 4 cos ACI4 8(B.1l,2)

where

B is given by (B.8,3).

(C)-!!:.. is the slope of the low-speed rolling moment due to yawing at zeroCL Ctt~g lift, obtained from Fig. B.ll,1 as a function of aspect ratio, sweep of

Page 359: Stability Control EtkinReid

s.u Wing Yawing Derivatives C1" c; 351

o 15

Ac/4 (degl

30 45 60

Taperratio

A.

/ I V V V 1.0/

V 1/ I 1/ /V ~/

I / £ V~- ~ -J --- --/- -/ --- k 0.5

1/ IV V /I

/ / V•...

I 0.25III

"' '-t--I

/ 1/ 1/ /v 1/ 1// V V 10-""'",./ I 0I

1// / /V V r :V V H....-~ I.-- I

I I

/VI V V V , I I I V t, I I I

/, I,' I

I

1//V V...V III IV III I I, 1/ II

I I / 1/" " t O~I/ f 4 6 f 11

0I I T .' , ,II I Aspect ratio, A

" " "I I II"1 I

III " " I

.1 .2 .3 .4 .5 .6

(~~)CL=OM=O

Figure B.ll,1 Wing yawing derivative C1r'

the quarter-chord, and taper ratio. (B.ll,2) modifies the low-speedvalue by means of the Prandtl-Glauert rule to yield approximate cor-rections for the first-order three-dimensional effects of compressibleflow up to the critical Mach number.

r

is the increment in CI, due to dihedral, given by

dCI, 1 7TA sin Acl4--=-r 12 A + 4 cos Acl4

is the geometric dihedral angle in radians, positive for the wing tipabove the plane of the root chord.

(B.ll,3)

is the increment in C1, due to wing twist obtained from Fig. B.ll ,2.

is the wing twist between the root and tip sections in degrees, nega-tive for washout (see Fig. B.ll,2).

The derivative Cn,:

C = ( Cn, ) C2 + ( Cn

, ) Cn, ci L C

DoDo

(B.ll,4)

Page 360: Stability Control EtkinReid

352 Appendix B. Data/or Estimating Aerodynamic Derivatives

.001

/L..--~

L..-- ./A.

~

L..-- po

/'

;{V'v.2 ~

I~ v-: ~/'

•.....

~V

VRoot-section

_ .J:ero-Iift line -~t~ <~-9 (deg~~~

-Tip-sectionzero-lift line

I I I I I2 8 10

.004

.003

AClr

9(per deg) .002

4 6Aspect ratio. A

Figure B.ll,2 Effect of wing twist on wing yawing derivative G1r

where

is the wing lift coefficient.

is the low-speed drag-due-to-lift yaw-damping parameter obtained fromFig. B.II,3 as a function of wing aspect ratio, taper ratio, sweepback, andCG position.

is the low-speed profile-drag yaw-damping parameter obtained from Fig.B.ll,4 as a function of the wing aspect ratio, sweep-back, and CG posi-tion.

is the wing profile drag coefficient evaluated at the appropriate Machnumber. For this application CDo is assumed to be the profile drag associ-ated with the theoretical ideal drag due to lift and is given by

where CD is the total drag coefficient at a given lift coefficient.

Page 361: Stability Control EtkinReid

B.12 Changes in Inertias and Stability Derivatives with Change of Body Axes 353

Taperatio

A.

1.0o

I IAc/4 (deg)

f'--. -60 (hnw -hi" 0

V!r!

~ t:::=:~ -~ c--: """"

~- ----- ---~~oy V : Ac/4 (hnw - h) " 0.2

1 0 1 : (deg)I

Cn/CLJ ~ I--.II

IV! ,1"-1 -50

/; I-40I

I .Q..I---I

V! ---:III

I,1 0 1 I

I

Cn/CL2\ Ac/4 (hnw - h) '" 0.4(deg)

"I -60:,,/; \1 <,r--

V! I\. r-~ -;} ...•••.-40I, J.---I 01---I

V! ,.vr,1 I 1 j 5 6

r i i 'tI I

1 0 1 Aspect ratio, A

Cn.,.ICL2

Figure B.ll,3 Low-speed drag-due-to-lift yaw-damping parameter.

Taperratio

A.

1.0o

Taperratio

A.

1.0o

B.12 Changes in Inertias and Stability Derivatives withChange of Body Axes

A matrix A that connects two vectors u and vas in (AA,23) transforms between tworeference frames as in (A.4,26). We now apply this rule to the inertia matrix and tomatrices of stability derivatives.

Page 362: Stability Control EtkinReid

354 Appendix B. Datafor Estimating Aerodynamic Derivatives

oo 2Aspect ratio, A

4 6 8 10I

Acl4la)(hnw-h) = 0(deg)

0~ -

~

40

V 50

60

-1

oo 2Aspect ratio, A

4 6 8 10

(bl (hn' -hI = 0'2 Ac/4w (deg)

0-'"~;;....- 40

V 50

V 60

.2-1

oo I IAc/4(e)(hnw-h) = 0.4(deg)

0.--~

40-Ii '/ 50

1//V 60

.2

2Aspect ratio. A4 6 8 10

-1

Figure B.ll,4 Low-speed profile-drag yaw-damping parameter.

Page 363: Stability Control EtkinReid

B.12 Changes in Inertias and Stability Derivatives with Change of Body Axes 355

TRANSFORMATION OF INERTIASThe inertia matrix I connects angular momentum with angular velocity [see (4.3,4)and (4.3,5)] via

h = Iw

and hence belongs to the class of matrices covered by (A.4,26). It follows that fortwo sets of body axes, denoted FBI and FB2 connected by the transformation L12, theinertias in frame FB2 can be obtained from those in FBI by

(B.12,1)

If the two frames are two sets of body axes such that xBI is rotated about YBI throughangle g to bring it to XB2' then (see Appendix A.4)

[

COS gO-sin g]L21 = 0 I 0

sin g 0 cos g(B.12,Z)

The inertias in frame FB2, denoted by an asterisk, are then obtained from those in FBI'

with the usual assumption of symmetry about the xz plane, by the relations

I: = I, cos" g + I, sin" g + Izx sin zgI; = l; sirr' g + I, cos" g - Izx sin 2g

-I;" = l(lx - Iz) sin 2g + Izx (sin? g - cos? g)(B.12,3)

TRANSFORMATION OF STABILITY DERIVATIVES

All of the stability derivatives with respect to linear and angular velocities and veloc-ity derivatives can be expressed as sums of expressions of the form of (A.4,23). Thatis, with the usual assumptions about separation of longitudinal and lateral motion, wecan write

[~]~p.0 X.fU] [0 Xq H~]+[~ 0 Of"]r, o v + Yp 0 0~w :wsz z; 0 z, aw 0 Zq 0

(B. 12,4)

[;]~ [~.Lv ° fU] [4 0

!]m+[~ 0

° f"]0 Mw v + 0 Mq 0~w :ws; o aw Np 0 0

(B.12,S)

Each of the six matrices of derivatives above transforms according to the rule(AA,26). When L is given by (B.12,2) we have the transformation from an initial setof body axes (unprimed) to a second set (primed) as follows:

Page 364: Stability Control EtkinReid

356 Appendix B. Data/or Estimating Aerodynamic Derivatives

Longitudinal

(Xu)' = Xu cos2 g - (Xw + Zu) sin g cos g + Z; sin2 g(Xw)' = Xw cos" g + (Xu - Zw) sin g cos g - Zu sirr' g(Xq)' = Xq cos g - Zq sin g(X,Y = Z", sin2 g (1)(X",)' = - Zw sin g cos € (1)

(Zu>' = Zu cos" g - (Z'w - Xu) sin g cos g - K; sin2 g(Zw)' = Zw cos? g + (Zu + Xw) sin g cos g + Xu sirr' g(Zq)' = Zq cos g + Xq sin g(Z,J' = -Z", sin g cos g (1)

(Z",)' = Z", cos" g(Mu>' = M; cos g - M; sin g(Mw)' = Mw cos g + M; sin g(Mq), = Mq

(M,Y = -M", sin g (1)

(M",)' = M", cos g (1)

Lateral

(Yv)' = Yv

(Yp)' = Yp cos g - Yr sin g(Yr)' = Yr cos g + Yp sin g(Lv)' = L; cos € - N; sin g(Lp)' = Lp cos" g - (L, + Np) sin € cos € + N, sin2 g(Lr)' = L; cos" g - (N, - Lp) sin ~ cos g - Np sin2 g(Nv)' = N; cos g + Lv sin g(Np)' = Np cos" ~ - (N, - Lp) sin € cos g - L; sin2 g(Nr)' = N; cos2 g + tL; + Np) sin € cos g + Lp sirr' €

(B.12,6)

(B. 12,?)

(1) For consistency of assumptions, the derivatives with respect to Ii and (X,,,)' are usually ignored.

Page 365: Stability Control EtkinReid

APPENDIX C

Mean Aerodynamic Chord,Mean Aerodynamic Center,

and Ci;acw

C.l Basic DefinitionsIn the normal flight range, the resultant aerodynamic forces acting on any lifting sur-face can be represented as a lift and drag acting at the mean aerodynamic center (x, y,z), together with a pitching couple Cmacw which is independent of angle of attack (seeFig. 2.8).

The pitching moment of a wing is nondimensionalized by the use of the meanaerodynamic chord c.

Both the m.a. center and the m.a. chord lie in the plane of symmetry of the wing.However, in determining them it is convenient to work with the half-wing.

These quantities are defined by (see Fig. C.I)!

2 Ib/2

c=- c2dy8 0

2 Ib/2

x= -8 CzcxdyCL 0 a

2 Ib/2 b

Y = CL8 0 Czacydy = 'Y/cP"2

2 Ib/2z = C

L8 0 CZacz dy

(C.I,I)

(C.1,2)

(C.I,3)

(C. 1,4)

where b = wing span

c = local chordCL = total lift coefficientCZa = local additional lift coefficient, proportional to CL

CZb= local basic lift coefficient, independent of CL

C/ = CZb + CZa = total local lift coefficient

lThe coordinate system used applies only to this appendix.

357

Page 366: Stability Control EtkinReid

358 Appendix C. Mean Aerodynamic Chord, Mean Aerodynamic Center and C\' mllCw

\

..-=-------------~y

Figure C.l Local aerodynamic center coordinates.

mac = pitching moment, per unit span, about aerodynamic center (Fig. C.4)S = wing areay = spanwise coordinate of local aerodynamic center measured from axis of

symmetryx = chordwise coordinate of local aerodynamic center measured aft of wing

apexz = vertical coordinate of local aerodynamic center measured from xy plane

TJcp = lateral position of the center of pressure of the additional load on the half-wing as a fraction of the semispan

The coordinates of the m.a. center depend on the additional load distribution;hence the position of the true m.a. center will vary with wing angle of attack if theform of the additional loading varies with angle of attack. For a wing that has noaerodynamic twist, the m.a. center of the half-wing is also the center of pressure ofthe half-wing. If there is a basic loading (i.e., at zero overall lift, due to wing twist),then (x, y, z) is the center of pressure of the additional loading.

The height and spanwise position of the local aerodynamic centers may be as-sumed known, and hence y and z for the half-wing can be calculated once the addi-tional spanwise loading distribution is known. However, in order to calculate x, thefore-and-aft position of each local aerodynamic center must be known first. If all thelocal aerodynamic centers are assumed to lie on the nth-chord line (assumed to bestraight), then

(C.l,5)

where c, = wing root chord

An = sweepback of nth-chord line, degrees

Ideal two-dimensional flow theory gives n = ~for subsonic speeds and n = t for su-personic speeds.

Page 367: Stability Control EtkinReid

C.2 Comparison of m.a. Chord and m.a. Center 359

The m.a. chord is located relative to the wing by the following procedure:

1. In (C.1,2) replace CIa by Ct» and for x use the coordinates of the i-chord line.2. The value of x so obtained (the mean quarter-chord point) is the ~-point of the

m.a. chord.

The above procedure and the definition of c (see C.I,I) are used for all wings.

C.2 Comparison of m.a. Chord and m.a. CenterforBasic Planforms and Loading Distributions

In Table C.I taken from (Yates, 1952), values of m.a. chord and yare given for somebasic p1anforms and loading distributions.

In the general case the additional loading distribution and the spanwise center-of-pressure position can be obtained by methods such as those of De Young and Harper(1948), Weissinger (1947), and Stanton-Jones (1950). For a trapezoidal wing with thelocal aerodynamic centers on the nth-chord line, the chordwise location of the meanaerodynamic center from the leading edge of the m.a. chord expressed as a fractionof the m.a. chord hnw is given by

3(1 + A)2 [ 1 + 2A ]hnw = n + 8(1 + A + A2) 'T/cp - 3(1 + A) A tan An (C.2,I)

Table cuAdditionalLoading MAC.

Planform Distribution C y

Constant taper and sweep Any 2cr 1 +,\ +,\2 b(trapezoidal) 3 1+'\ 'T/cp' "2

Constant taper and sweep Proportional 2cr 1 +,\ +,\2 b 1 + 2,\(trapezoidal) to wing chord 3 1+'\ 2 3(1 + ,\)

(uniform CI)

Constant taper and sweep Elliptic 2cr 1 +,\ +,\2 b 4(trapezoidal) 3 1+'\ 2 3'7T

Elliptic (with straight Any c, 8 bsweep of line of local a.c.) 3 '7T 'T/cp' "2

Elliptic (with straight Elliptic (uniform c, 8 b 4sweep of line of local a.c.) CI) 3 '7T 2 3'7T

Any (with straight sweep of Elliptic 2 Ib/2 b 4line oflocal a.c.) S 0 ~dy 2 3'7T

Page 368: Stability Control EtkinReid

360 Appendix C. Mean Aerodynamic Chord, Mean Aerodynamic Center and Cm""w

whereA = aspect ratio, b2/S

A = taper ratio, c/cr

c, = wing-tip chord

The length of the chord through the centroid of area of a trapezoidal half-wing isequal to c. For the same wing with uniform spanwise lift distribution (i.e., Cia =const) and local aerodynamic centers on the nth-chord line, the m.a. center also lieson the chord through the centroid of area. The chord through the centroid of area of awing having an elliptic planform is not the same as c, but the m.a, center for ellipticloading and the centroid of area both lie on the same chord (see Yates, 1952).

C.3 m.a. Chord and m.a. Centerfor Swept and TaperedWings (Subsonic)

The ratio cler is plotted against A in Fig. C.2 for straight tapered wings with stream-wise tips. The spanwise position of the m.a. center of the half-wing (or the center ofpressure of the additional load) for uniform spanwise loading is also given in Fig.C.2. These functions are given in Table Ci l.

The m.a. chord is located by means of the distance x of the leading edge of them.a. chord aft of the wing apex:

b 1 1 + 2Ax=2"'3 l+A tan A,

1 + 2A= 12 crA tan Ao (C.3,1)

where Ao = sweepback of wing leading edge, degrees.The sweepback of the leading edge is related to the sweep of the nth-chord line

An by the relation

I-AAtanAo=AtanAn+4n 1 +A (C.3,2)

Using (C.3,2) and the expression for cle,., x can be obtained in terms of c and An from

x (1 + 2A)(1 + A) [ 1 - A ]- = AtanA +4n--c 8(1 + A + A2) n 1 + A

(C.3,3)

The fractional distance of the m.a. center aft of the leading edge of the m.a.chord, hnw' is given for swept and tapered wings at low speeds and small incidencesin Fig. C.3. The dotted lines show the aerodynamic-center position for wings withunswept trailing edges. The curves have been obtained from theoretical and experi-mental data. The curves apply only within the linear range of the curve of wing liftagainst pitching moment, provided that the flow is subsonic over the entire wing. Theprobable error of hnw given by the curves is within 3%.

Page 369: Stability Control EtkinReid

1.0

0.9

1 •• 1 ••••

0.8

0.7

0.6

0.5

I~~

0.4

vC 21+A+A2 /C;=3" 1+A ~

//

----'For uniformspanwise loading

i 1 + 2Ab/2 = 3(1 + AIl>-V---

~

V0.3 o 0.2 0.4 0.6 0.8 1.0

Figure C.2 Mean aerodynamic chord for straight tapered wings; and spanwise position of meanaerodynamic center for uniform spanwise loading (i.e., constant CL). (From "Notes on the MeanAerodynamic Chord and the Mean Aerodynamic Center of a Wing" by A. H. Yates, J. Roy. Aero.Soc., June 1952.)

C.4The total load on each section of a wing has three parts as illustrated by Fig. C.4a.The resultant of the local additional lift la' is the lift La acting through the m.a. center(Fig. CAb).

The resultant of the distribution of the local basic lift l» is a pitching couplewhenever the line of aerodynamic centers is not straight and perpendicular to x. Thiscouple is given by

Page 370: Stability Control EtkinReid

30 401\1/4' deg

Figure C.3 Chordwise position of the mean aerodynamic center of swept and tapered wings atlow speeds expressed as a fraction of the mean aerodynamic chord. (From Royal Aeronautical DataSheet Wings 08.01.01.)

362 Appendix C. Mean Aerodynamic Chord, Mean Aerodynamic Center and Cm""w

0.6.-----:--------,r----..,.----.------.------.

0.5

0.41----.---.....:..:...,.....::....---1----

0.3t---t---t~7_F_-___::::~~-+-~4_--__l

10 30 40/\'/4' deg

50 6020

II.l

A=0.5 .•.~/ ./---~~V- -•..._- '.0 Cropped- _ -./ delta wings_._~O.S _

>

0.4

0.3

0.2

oo 5030 40/\1/4' deg

6010 20

II.,f

~A=6.04.0I--... h~iO"'-- -

••....1~0 ---~

J.5 -L--

A=11.0

0.3

0.2

0.1

10 50 6020

70

70

70

Page 371: Stability Control EtkinReid

(al

(bl

Figure C.4 (a) Section total load. (b) Wing loads.

since the resultant of lb = O.Then

4 Lb/2 1

c.; = py'2Sc 0 »c; 2" pV2c dy

2 Ib12

= Sc 0 C1bxc dy

The resultant of the mac distribution is given by

2 Ib12

Cm2 = Sc 0 Cmacc2 dy

The total pitching-moment coefficient about the m.a. center is then

(CA,l)

(CA,2)

(CA,3)

If Cmac is constant across the span, and equals Cm2, then (CA,2) also becomes thedefining equation for C.

Page 372: Stability Control EtkinReid

APPENDIX D

The Standard Atmosphereand Other Data

The Standard Atmosphere

The tables that follow are derived from The ARDC Model Atmosphere, 1959, byMinzner, R. A., Champion, K. S. W., and Pond, H. L. Air Force Cambridge Research ••Center Report No. TR-59-267, U.S. Air Force, Bedford, MA, 1959. The values in thetables are the same, for most engineering purposes, as those derived from U.S. STAN-DARD ATMOSPHERE, 1976. Prepared by the USAF, NASA, and the NOAA.

English Units"

Speed of KinematicAltitude Temperature Pressure P, Density p, sound, viscosity,

h,ft T, OR lb/ff lb seCZ!ft4 ft/sec tf/sec

0 518.69 2116.2 2.3769-3 1116.4 1.572r4

1,000 515.12 2040.9 2.3081 1112.6 1.61052,000 511.56 1967.7 2.2409 1108.7 1.64993,000 507.99 1896.7 2.1752 1104.9 1.69054,000 504.43 1827.7 2.1110 1101.0 1.73245,000 500.86 1760.9 2.0482 1097.1 1.7755

6,000 497.30 1696.0 1.9869-3 1093.2 1.8201-4

7,000 493.73 1633.1 1.9270 1089.3 1.86618,000 490.17 1572.1 1.8685 1085.3 1.91369,000 486.61 1512.9 1.8113 1081.4 1.9626

10,000 483.04 1455.6 1.7556 1077.4 2.0132

11,000 479.48 1400.0 1.7011-3 1073.4 2.0655-4

12,000 475.92 1346.2 1.6480 1069.4 2.119613,000 472.36 1294.1 1.5961 1065.4 2.175414,000 468.80 1243.6 1.5455 1061.4 2.233115,000 465.23 1194.8 1.4962 1057.4 2.2927

16,000 461.67 1147.5 1.4480-3 1053.3 2.3544-4

17,000 458.11 1101.7 1.4011 1049.2 2.418318,000 454.55 1057.5 1.3553 1045.1 2.484319,000 450.99 1014.7 1.3107 1041.0 2.552620,000 447.43 973.27 1.2673 1036.9 2.6234

364

Page 373: Stability Control EtkinReid

The Standard Atmosphere 365

English Units" (Continued)

Speed of Kinematic

Altitude Temperature Pressure P, Density p, sound, viscosity,

h.ft T. -s lb/ff lb sec2/ft4 ft/sec ff/sec

21,000 443.87 933.26 1.2249-3 1032.8 2.6966-4

22,000 440.32 894.59 1.1836 1028.6 2.7724

23,000 436.76 857.24 1.1435 1024.5 2.8510

24,000 433.20 821.16 1.1043 1020.3 2.9324

25,000 429.64 786.33 1.0663 1016.1 3.0168

26,000 426.08 752.71 1.0292-3 1011.9 3.1044-4

27,000 422.53 720.26 9.9311-4 1007.7 3.1951

28,000 418.97 688.96 9.5801 1003.4 3.2893

29,000 415.41 658.77 9.2387 000.13 3.3870

30,000 411.86 629.66 8.9068 994.85 3.4884

31,000 408.30 601.61 8.5841-4 990.54 3.593T4

32,000 404.75 574.58 8.2704 986.22 3.703033,000 401.19 548.54 7.9656 981.88 3.8167

34,000 397.64 523.47 7.6696 977.52 3.9348

35,000 394.08 499.34 7.3820 973.14 4.0575

36,000 390.53 476.12 7.1028-4 968.75 4.1852-4

37,000 389.99 453.86 6.7800 968.08 4.3794

38,000 389.99 432.63 6.4629 968.08 4.5942

39,000 389.99 412.41 6.1608 968.08 4.819640,000 389.99 393.12 5.8727 968.08 5.0560

41,000 389.99 374.75 5.5982-4 968.08 5.3039-4

42,000 389.99 357.23 5.3365 968.08 5.5640

43,000 389.99 340.53 5.0871 968.08 5.836844,000 389.99 324.62 4.8493 968.08 6.123045,000 389.99 309.45 4.6227 968.08 6.4231

46,000 389.99 294.99 4.406T4 968.08 6.7380-4

47,000 389.99 281.20 4.2008 968.08 7.068248,000 389.99 268.07 4.0045 968.08 7.414649,000 389.99 255.54 3.8175 968.08 7.778050,000 389.99 243.61 3.6391 968.08 8.1591

51,000 389.99 232.23 3.4692-4 968.08 8.5588-4

52,000 389.99 221.38 3.3072 968.08 8.978153,000 389.99 211.05 3.1527 968.08 9.417954,000 389.99 201.19 3.0055 968.08 9.879255,000 389.99 191.80 2.8652 968.08 1.036r3

56,000 389.99 182.84 2.7314-4 968.08 1.0871-3

57,000 389.99 174.31 2.6039 968.08 1.140358,000 389.99 166.17 2.4824 968.08 1.196159,000 389.99 158.42 2.3665 968.08 1.254760,000 389.99 151.03 2.2561 968.08 1.3161

Page 374: Stability Control EtkinReid

366 Appendix D. The Standard Atmosphere and Other Datil

English Units" (Continued)

Speed of KinematicAltitude Temperature Pressure P, Density p, sound, viscosity,

h,ft T, oR lbljf lb sec2/fi4 ft/sec ~/sec

61,000 389.99 143.98 2.1508-4 968.08 1.3805-3

62,000 389.99 137.26 2.0505 968.08 1.448163,000 389.99 130.86 1.9548 968.08 1.518964,000 389.99 124.75 1.8636 968.08 1.593265,000 389.99 118.93 1.7767 968.08 1.6712

66,000 389.99 113.39 1.6938-4 968.08 1.7530-3

67,000 389.99 108.10 1.6148 968.08 1.838768,000 389.99 102.06 1.5395 968.08 1.928669,000 389.99 98.253 1.4678 968.08 2.023070,000 389.99 93.672 1.3993 968.08 2.1219

71,000 389.99 89.305 1.3341-4 968.08 2.225T3

72,000 389.99 85.142 1.2719 968.08 2.334573,000 389.99 81.174 1.2126 968.08 2.448674,000 389.99 77.390 1.1561 968.08 2.568375,000 389.99 73.784 1.1022 968.08 2.6938

"Note: the notation =-0 means = X 10-0. ..

SI Units

Tempera- Speed of KinematicAltitude ture Pressure Density Sound Viscosity,

h,m T,K PN/m2 p kg/m3 m/s m2/s ",

0 288.16 1.01325+5 1.2250 340.29 1.4607-5

300 286.21 9.7773+4 1.1901 339.14 1.4956600 284.26 9.4322 1.1560 337.98 1.5316900 282.31 9.0971 1.1226 336.82 1.5687

1,200 280.36 8.7718 1.0900 335.66 1.60691,500 278.41 8.4560 1.0581 334.49 1.6463

1,800 276.46 8.1494 1.0269 333.32 1.6869

2,100 274.51 7.8520 9.9649-1 332.14 1.7289

2,400 272.57 7.5634 9.6673 330.96 1.7721

2,700 270.62 7.2835 9.3765 329.77 1.8167

3,000 268.67 7.0121 9.0926 328.58 1.86283,300 266.72 6.7489 8.8153 327.39 1.9104

3,600 264.77 6.4939 8.5445 326.19 1.9595

3,900 262.83 6.2467 8.2802 324.99 2.01024,200 260.88 6.0072 8.0222 323.78 2.06264,500 258.93 5.7752 7.7704 322.57 2.11674,800 256.98 5.5506 7.5247 321.36 2.17275,100 255.04 5.3331 7.2851 320.14 2.2305

5,400 253.09 5.1226 7.0513 318.91 2.29035,700 251.14 4.9188+4 6.8234-1 317.69 2.352Z-5

Page 375: Stability Control EtkinReid

The Standard Atmosphere 367

SI Units (Continued)

Tempera- Speed of KinematicAltitude ture Pressure Density Sound Viscosity,

h,m T,K PN/m2 pkg/m3 m/s m2/s

6,000 249.20 4.7217 6.6011 316.45 2.41616,300 247.25 4.5311 6.3845 315.21 2.48246,600 245.30 4.3468 6.1733 313.97 2.55096,900 243.36 4.1686 5.9676 312.72 2.62187,200 241.41 3.9963 5.7671 311.47 2.69537,500 239.47 3.8299 5.5719 310.21 2.77147,800 237.52 3.6692 5.3818 308.95 2.85038,100 235.58 3.5140 5.1967 307.68 2.93208,400 233.63 3.3642 5.0165 306.41 3.01678,700 231.69 3.2196 4.8412 305.13 3.10469,000 229.74 3.0800 4.6706 303.85 3.19579,300 227.80 2.9455 4.5047 302.56 3.29039,600 225.85 2.8157 4.3433 301.27 3.38849,900 223.91 2.6906 4.1864 299.97 3.4903

10,200 221.97 2.5701 4.0339 298.66 3.596110,500 220.02 2.4540 3.8857 297.35 3.706010,800 218.08 2.3422 3.7417 296.03 3.820211,100 216.66 2.2346 3.5932 295.07 3.956411,400 216.66 2.1317 3.4277 295.07 4.147411,700 216.66 2.0335+4 3.2699-1 295.07 4.3475-5

12,000 216.66 1.9399 3.1194 295.07 4.557412,300 216.66 1.8506 2.9758 295.07 4.777312,600 216.66 1.7654 2.8388 295.07 5.007812,900 216.66 1.6842 2.7081 295.07 5.249413,200 216.66 1.6067 2.5835 295.07 5.502613,500 216.66 1.5327 2.4646 295.Q7 5.768013,800 216.66 1.4622 2.3512 295.Q7 6.046214,100 216.66 1.3950 2.2430 295.07 6.337814,400 216.66 1.3308 2.1399 295.Q7 6.643414,700 216.66 1.2696 2.0414 295.07 6.963715,000 216.66 1.2112 1.9475 295.Q7 7.299515,300 216.66 1.1555 1.8580 295.07 7.651415,600 216.66 1.1023 1.7725 295.07 8.020215,900 216.66 1.0516 1.6910 295.07 8.406816,200 216.66 1.0033 1.6133 295.07 8.811916,500 > 216.66 9.5717+3 1.5391 295.Q7 9.236616,800 216.66 9.1317 1.4683 295.07 9.681617,100 216.66 8.7119 1.4009 295.07 1.0148-4

17,400 216.66 8.3115 1.3365 295.07 1.063717,700 216.66 7.9295+3 1.2751-1 295.07 1.1149-418,000 216.66 7.5652 1.2165 295.07 1.168618,300 216.66 7.2175 1.1606 295.07 1.224918,600 216.66 6.8859 1.1072 295.07 1.283918,900 216.66 6.5696 1.0564 295.07 1.3457

Page 376: Stability Control EtkinReid

368 Appendix D. The Standard Atmosphere and Other Data

Other DataConversion Factors

Multiply By TaGet

Pounds (lb)Feet (ft)SlugsSlugs per cubic foot (slugs/ft")Miles per hour (mph)Knots (kt)Knots (kt)

4.4480.304814.59515.40.44710.51511.152

Newtons (N)Meters (m)Kilograms (kg)Kilograms per cubic meter (kg/m3)

Meters per second (mls)Meters per second (mls)Miles per hour (mph)

Gravityg = 32.2 ftls2 = 9.81 m/s2 at sea level.

Page 377: Stability Control EtkinReid

APPENDIX E

Datafor the Boeing 747-100

The Boeing 747 is a highly successful, large, four-engined turbofan transport aircraft.The model 100 first entered service in January 1970, and since then it has continuedto be developed through a series of models and special versions. As of May 1990,only versions of the model 400 were being marketed. By the year 1994, close to 800Boeing 747s were in operation around the world, and the aircraft was still in produc-tion.

The data for the Boeing 747-100 contained in this appendix are based on Heffleyand Jewell (1972). A three-view drawing of the aircraft is given in Fig. E.l. A bodyaxis system FB is located with origin at the CG and its x-axis along the fuselage refer-ence line (FRL). The CG is located at 0.25 c (i.e., h = 0.25), and this is the locationthat applies for the tabulated data. The thrust line (TL) makes an angle of 2.50 withrespect to the FRL as shown.

Three flight cases are documented in the data tables. They all represent straightand level steady-state flight at a fixed altitude. Case I has the aircraft in its landingconfiguration with 300 flaps, landing gear down, and an airspeed 20% above thestalling speed. Cases IT and 1lI represent two cruising states with the flaps retractedand the gear up.

The data in Table E.1 define the flight conditions that apply to the three cases. Itshould be noted that the moments and product of inertia are given relative to the bodyframe FB shown in Fig. E.l. Here the weight and inertias for Case I are smaller thanthose for the other two cases because the amount of fuel on board during landing isless than that during the cruise. If the data are to be applied to a reference frame dif-ferent from FB (e.g., to stability axes Fs) then the given inertias will have to be trans-formed according to (B.12,3). Note that FB can be rotated into Fs by a single rotationof g about the y-axis. Values for g are contained in Table E.1.

The dimensional derivatives corresponding to FB of Fig. E.1 are contained in Ta-bles E.2 to EA. Since FB can be rotated into the stability axes Fs by a single rotationof g about the y-axis, it follows that the transformations of (B.12,6 and B.12,7) can beused to obtain the derivatives corresponding to Fs- Values for g are contained in TableE.1.

369

Page 378: Stability Control EtkinReid

...Ja: ...Ju.. I-

oo

o

370

Page 379: Stability Control EtkinReid

Appendix E. Data/or the Boeing 747-100 371

TableE.1Boeing 747-100 Data

(S = 5,500 ff, b = 195.68 ft, C = 27.31 ft, h = 0.25)

Case I Case II Case III

Altitude (ft) 0 20,000 40,000

M 0.2 0.5 0.9

V (ftIs) 221 518 871

W(1b) 5.640 X lOs 6.366 X lOS 6.366 X lOS

I; (slug-ft") 1.42 X 107 1.82 X 107 1.82 X 107

Iy (slug-fr') 3.23 X 107 3.31 X 107 3.31 X 107

I, (slug-fr') 4.54 X 107 4.97 X 107 4.97 X 107

Iv< (slug-ff') 8.70 X lOs 9.70 X lOS 9.70 X lOS

~(degrees) -8.5 -6.8 -2.4

CD 0.263 0.040 0.043

TableE.2Boeing 747-100 Dimensional Derivatives

Case I (M = 0.2)Longitudinal

X (Zb) Z(Zb) M (fl·Zb)

u (ftls) -3.661 X 102 -3.538 X 1Q3 3.779 X 103

w (ftIs) 2.137 X 103 -8.969 X 103 -5.717 X 104

q (rad/s) 0 -1.090 X lOs -1.153 X 107

W(ft/s") 0 5.851 X 102 -7.946 X 103

B. (rad) 1.680 X 1if -1.125 X lOS -1.221 X 107

Lateral

Y(Zb) L (fl'Zb) N(fl·Zb)

v (ft/s) -1.559 X 103 -8.612 X 1if 3.975 X 104

p (rad/s) 0 -1.370 X 107 -6.688 X 106

r (rad/s) 0 4.832 X 106 -1.014 X 107

s: (rad) 0 -3.200 X 106 -1.001 X 106

Br (rad) 5.729 X 104 1.034 X 106 -6.911 X 106

Page 380: Stability Control EtkinReid

372 Appendix E. Data for the Boeing 747-100

TableE.3Boeing 747-100 Dimensional Derivatives

Case II (M = 0.5)Longitudinal

X (lb) Z(lb) M (ft-lb)

u (fils) -4.883 X 101 -1.342 X 103 8.176 X 103

w (fils) 1.546 X 103 -8.561 X 103 -5.627 X 104

q (rad/s) 0 -1.263 X UP -1.394 X 107

W (fi/s2) 0 3.104 X 1Q2 -4.138 X 103

8. (rad) 3.994 X 1if -3.341 X 105 -3.608 X 107

Lateral

Y(lb) L (fl'lb) N(fl·lb)

v (fils) -1.625 X 103 -7.281 X 104 4.404 X 1Q4

p (rad/s) 0 -1.180 X 107 -2.852 X 106

r (rad/s) 0 6.979 X 106 -7.323 X 106

8a (rad) 0 -2.312 X 106 -7.555 X lOS

s, (rad) 1.342 X lOS 3.073 X 106 -1.958 X 107

\

TableE.4Boeing 747-100 Dimensional Derivatives

Case ill (M = 0.9)Longitudinal

X(lb) Z(lb) M (fl·lb)

u (fils) -3.954 X 1Q2 -8.383 X 1Q2 -2.062 X 103

w (ftIs) 3.144 X 1Q2 -7.928 X 103 -6.289 X 104

q (rad/s) 0 -1.327 X 105 -1.327 X 107

W(fi/s2) 0 1.214 X 102 -5.296 X 103

8. (rad) 1.544 X 1if -3.677 X 105 -4.038 X 107

Lateral

Y(lb) L (fl·lb) N (ft'lb)

v (fils) -1.198 X 103 -2.866 X 104 5.688 X 104

p (rad/s) 0 -8.357 X 106 -5.864 X 105

r (rad/s) 0 5.233 X 106 -7.279 X 106

s, (rad) 0 -3.391 X 106 4.841 X lOS

s, (rad) 7.990 X 104 2.249 X 106 -2.206 X 107

Page 381: Stability Control EtkinReid

References 373

ReferencesReferences CitedAGARD. Aeroelastic Effects from a Flight Mechanics Standpoint. Conference Proceedings, no. 46,

March 1970.AGARD. Flight Test Manual, Vol. II, Stability and Control. AGARD, 1959.AGARD. Rotorcraft System Identification. Advisory Report No. 280, NATO, 1991.ANSI/AIAA. Recommended Practise-Atmospheric and Space Flight Vehicle Coordinate Sys-

tems. AIAA, Washington, DC, 1992.Bairstow, L. AppliedAerodynamics, 2nd ed. Longmans, Green & Co., 1939.Bisplinghoff, Raymond L. Principles of Aeroelasticity. John Wiley & Sons, Inc., New York, 1962.Bollay, W. Aerodynamic Stability and Automatic Control. Jour. Aero. Sci. vol. 18, No.9, 1951.Bryan, G. H. Stability in Aviation. Macmillan Co., London, 1911.Bur. Aero. The Human Pilot. Vol. ill of Bur. Aero. Rept. AF-6-4, 1954.Byrne, J. E. Design of a Lateral Ride Comfort Control System for STOL Aircraft. University of

Toronto UTIAS Tech. Note No. 247, 1983.Campbell, J. P., and McKinney, M. O. Summary of Methods for Calculating Dynamic Lateral Sta-

bility and Response and for Estimating Lateral Stability Derivatives. NACA Rept. 1098, 1952.Carslaw, H. S., and Jaeger, J. C. Operational Methods in Applied Mathematics, 2nd ed. Oxford

University Press, London, 1947.Cooper, G. E., and Harper, R. P., Jr. The Use of Pilot Rating in the Evaluation of Aircraft Handling

Qualities. NASA TN D-5153, 1969.Davenport, A. G. Rationale for Determining Design Wind Velocities. J. Struct. Div. ASCE, vol. 86,

1960.De Young, J., and Harper, C. W. Theoretical Symmetric Span Loadings at Subsonic Speeds for

Wings Having Arbitrary Planforms. NACA Rept. 921,1948.Dowell, Earl H., et al., ed. A Modem Course in Aeroelasticity. Solid Mechanics and Its Applica-

tions, vol. 32. Dordrecht, Boston: Kluwer Academic, 1994.Duncan, W. J. Control and Stability of Aircraft. Cambridge University Press, Cambridge, 1952.Ellis, N. D. Aerodynamics of Wing-Slipstream Interaction. Univ. of Toronto UTIAS Rept. 169,

1971.Etkin, B. Aerodynamic Transfer Functions: An Improvement on Stability Derivatives for Unsteady

Flight. Univ. of Toronto UTIA Rept. 42, 1956.Etkin, B. Dynamics of Flight-Stability and Control. 1st ed., John Wiley & Sons, Inc., New York,

1959.Etkin, B., and Hughes, P. C. Explanation of the Anomalous Spin Decay of Satellites with Long

Flexible Antennae. J. Spacecraft, vol. 4, no. 9, Sept. 1967.Etkin, B. Dynamics of Atmospheric Flight. John Wiley & Sons, Inc., New York, 1972.Etkin, B. Turbulent Wind and Its Effect on Flight. AIAA Jour. of Aircraft, vol. 18, no. 5,1981.Etkin, B. Dynamics of Flight, Stability and Control. 2nd ed. John Wiley & Sons, Inc., New York,

1982.Etkin, B., and Etkin, D. A. Critical Aspects of Trajectory Prediction: Flight in Non-uniform Wind,

in AGARDograph No. 301, v. 1 (1990).Filotas, L. T. Approximate Transfer Functions for Large Aspect Ratio Wings in Turbulent Flow. J.

Aircraft, vol. 8, no. 6, June 1971.Flax, A. H., and Lawrence, H. R. The Aerodynamics of Low-Aspect Ratio Wing-Body Combina-

tions. ComellAeroLab. Rept. CAL-37, 1951.Fung, Y. C. The Theory of Aeroelasticity. John Wiley & Sons, Inc., New York, 1955.Gates, S. B. Proposal for an Elevator Manoeuvreability Criterion. ARC R & M 2677, 1942.Heffley, R. K., and Jewell, N. F. Aircraft Handling Qualities Data. NASA CR 2144, Dec. 1972.Heppe, R. R., and Celinker, L. Airplane Design Implications of the Inertia Coupling Problem. lAS

Preprint 723, 1957.Hughes, P. C. Spacecraft Attitude Dynamics. John Wiley & Sons, Inc., New York, 1986.ICAO. Airworthiness of Aircraft. Anex 8, Chap. 2; para 2.3; Flying Qualities, 1991.

Page 382: Stability Control EtkinReid

374 Appendix E. Data for the Boeing 747·100

Jaeger, J. C. An Introduction to the Laplace Transformation. Methuen & Co., London, 1949.Kleinman, D. L., Baron, S., and Levison, W. H. An Optimal Control Model of Human Response.

Part I: Theory and Validation. Automatica; vol. 6, pp. 357-369,1970.Kuethe, A. M., and Chow, C. Y. Foundations of Aerodynamics. 3rd ed., John Wiley & Sons, Inc.,

New York, 1976.Kuhn, R. E. Semiempirical Procedure for Estimating Lift and Drag Characteristics of Propeller-

Wing-Flap Configurations for V/STOL Airplanes. NASA Memo 1-16-59L, Feb. 1959.Lanchester, F. W. Aerodonetics. A. Constable & Co. Ltd., London, 1908.Langhaar, H. L. Dimensional Analysis and Theory of Models. John Wiley & Sons, Inc., New York,

1951.Maine, R. E., and lliff, K. W. Application of Parameter Estimation to Aircraft Stability and Con-

trol. NASA Reference Publication no. 1168, 1986.Margason, R. J., Hammond, A. D., and Gentry, G. L. Longitudinal Stability and Control Character-

istics of a Powered Model of a Twin-Propeller Deflected-Slipstream STOL Airplane Configu-ration. NASA TN D-3438, 1966.

McLean, D. Automatic Flight Control Systems. Prentice Hall International (U.K.) Ltd., 1990.McCormick, B. W. Aerodynamics, Aeronautics, and Flight Mechanics. John Wiley & Sons, Inc.,

New York, 1994.McLaughlin, M. D. A Theoretical Investigation of the Short-Period Dynamic Longitudinal Stability

of Airplanes Having Elastic Wings of 0° to 60° Sweepback. NACA TN 3251, 1954.McRuer, D. T., and Krendel, E. S. Mathematical Models of Human Pilot Behavior. AGARDograph

M-188, Nov. 1973.McRuer, D., Ashkenas, I., and Graham, D. Aircraft Dynamics and Automatic Control. Princeton

Univ. Press, Princeton, NJ, 1973.Michael, W. H., Jr. Analysis of the Effects of Wing Interference on the Tail Contributions to the

Rolling Derivatives. NACA Rept. 1086, 1952.Miele, A. Flight Mechanics-l Theory of Flight Paths. Addison-Wesley Pub. Co., Reading, MA,

1962.Miles, J. W. On the Compressibility Correction for Subsonic Unsteady Flow. J. Aero. Sci., vol. 17,

no. 3,1950.Miles, J. W. Unsteady Flow Theory in Dynamic Stability. J. Aero. Sci., vol. 17, no. 1, 1950.Millikan, C. B. The Influence of Running Propellors on Airplane Characteristics. J. Aero Sci., vol.

7, no. 3, 1940.Milne, R. D. Dynamics of the Deformable Airplane. ARC R & M 3345, 1964.Neumark, S. Longitudinal Stability, Speed and Height. RAE Rept. Aero 2265, 1948; also Aircraft

Engineering, Nov. 1950.Nyquist, H. Regeneration Theory. Bell System Tech. Jour. vol. 11, Jan. 1932.O'Hara, F. Handling Criteria. J. Roy. Aero. Soc., vol. 71, no. 676, pp. 271-291, 1967.Perkins, C. D. Development of Airplane Stability and Control Technology. J. Aircraft, vol. 7, no. 4,

1970.Phillips, W. H. Effect of Steady Rolling on Longitudinal and Directional Stability. NACA TN. 1627,

1948.Pinsker, W. J. G. Charts of Peak Amplitudes in Incidence and Sideslip in Rolling Manoeuvres Due

to Inertia Cross Coupling. RAE Rept. No. Aero 2604, April 1958.Priestley, E. A General Treatment of Static Longitudinal Stability with Propellors, with Application

to Single Engine Aircraft. ARC R & M 2732, 1953.Reid, L. D. The Measurement of Human Pilot Dynamics in a Pursuit-Plus-Disturbance Task. Univ.

of Toronto UTIAS Rept. 138, 1969.Ribner, H. S. Notes on the Propellor and Slip-Stream in Relation to Stability. NACA Wartime Rept.

U5,1944.Ribner, H. S. Formulas for Propellors in Yaw and Charts of the Side-Force Derivative. NACA Rept.

819,1945.Ribner, H. S. Field of Flow About a Jet and Effect of Jets on Stability of Jet-Propelled Airplanes.

NACA Wartime Rept. Ul3, 1946.

Page 383: Stability Control EtkinReid

References 375

Ribner, H. S., and Ellis, N. D. Aerodynamics of Wing-slipstream Interaction. CASI Transactions,vol. 5, no. 2, 1972.

Rodden, W. P. An Aeroelastic Parameter for Estimation of the Effect of Flexibility on the LateralStability and Control of Aircraft. J. Aero. Sci., vol. 23, no. 7, 1956.

Rodden, W. P., and Giesing, J. P. Application of Oscillatory Aerodynamic Theory to Estimation ofDynamic Stability Derivatives. J. Aircraft, vol. 7, no. 3, 1970.

Routh, E. J. Dynamics of a System of Rigid Bodies, 6th ed. Macmillan & Co., London, 1905.Rynaski, E. G., Andrisani, D. III, Weingarten, N. Active Control for the Total-in-Flight Simulator.

NASA CR 31I8, 1979.Schlichting, H., and Truckenbrodt, E. Aerodynamics of the Airplane. McGraw Hill International

Book Co., New York, 1979.Smelt, R, and Davies, H. Estimation of Increase in Lift Due to Slipstream. ARC R & M 1788,

1937.Stanton-Jones, R An Empirical Method for Rapidly Determining the Loading Distributions on

Swept Back Wings. ColI. Aero. (Cranfield) Rept. 32, 1950.Stevens, B. L., and Lewis, F. L. Aircraft Control and Simulation. John Wiley and Sons, Inc., New

York,1992.Synge, J. L., and Griffith, B. A. Principles of Mechanics. McGraw-Hill Pub. Co., Inc., New York,

1942.Theodorsen, Th. General Theory of Aerodynamic Instability and the Mechanism of Flutter. NACA

Rept.496, 1934.Tobak, M. On the Use of the Indicial Function Concept in the Analysis of Unsteady Motions of

Wings and Wing-Tail Combinations. NACA Rept. 1188, 1954.Truitt, R W. Hypersonic Aerodynamics. Ronald Press, New York, 1959.USAF. Stability and Control Datcom. Air Force Flight Dynamics Lab., U.S. Air Force, 1960 (re-

vised 1978).USAF. Flying Qualities of Piloted Airplanes. USAF Spec. MIL-F-8785C, 1980.USAF. Flying Qualities of Piloted Vehicles. USAF MIL-STD-1797 A, 1990.Weil, J., and Sleeman, W. c., Jr. Prediction of the Effects of Propellor Operation on the Static Lon-

gitudinal Stability of Single-Engine Tractor Monoplanes with Flaps Retracted. NACA Rept.941, 1949.

Weissinger, 1. The Lift Distribution of Swept Back Wings. NACA TM 1120, 1947.Yates, A. H. Notes on the Mean Aerodynamic Chord and the Mean Aerodynamic Centre of a

Wing. J. Roy. Aero. Soc., vol. 56, p. 461, June 1952.

Additional ReadingAnderson, J. D. Introduction to Flight. 2nd ed. McGraw Hill, Inc., New York, 1985.Bertin, J. J., and Smith, M. 1. Aerodynamicsfor Engineers. 2nd ed. Prentice Hall, NJ, 1989.Blakelock, J. H. Automatic Control of Aircraft and Missiles. 2nd ed. John Wiley & Sons, Inc., New

York,1991.McCormick, B. W. Aerodynamics, Aeronautics, and Flight Mechanics. John Wiley & Sons, Inc.,

New York, 1994.Nelson, R C. Flight Stability and Automatic Control. McGraw Hill Inc., New York, 1989.Shevell, R S. Fundamentals of Flight. 2nd ed. Prentice Hall, NJ, 1989.

Page 384: Stability Control EtkinReid
Page 385: Stability Control EtkinReid

IN D E X

Accelerometer, 271Adjoint matrix, 209, 304Aeroballistic range, 6Aerodynamic center, 24, 356Aerodynamic transfer functions, III, 129Aeroelastic effects, 29Aeroelastic oscillations, 5AFCS, 7, 206Aileron, 86

adverse yaw, 87reversal, 72, 88

Airspeed,15Altitude and glide path control, 275Angle of attack, 17Angle of sideslip, 17Angular momentum, 95, 96Apparent mass, 145Approximation:

Dutch Roll mode, 195,251longitudinal modes, 171roll mode, 193spiral mode, 193

Atmosphere standard, 363Atmospheric turbulence, response to, 106, 295Automatic flight control system(AFCS), 7,

206Autopilot, 8Autorotation, 151Axes, 15, 101, 102

Bairstow,3Bernoulli, 3Bisplinghoff, 73Bode diagram, 217, 227Body axes, 16, 101, 102Boeing 747, data, 368Boost gearing, 49Bryan, 3, 110Buckingham's 'IT theorem, 115Buffeting, 72

Canard configuration, 23Centre of pressure, 134CG limits, 74Characteristic:

determinant, 162equation, 162polynomial, 162,209

Cofactor, 304Complex amplitude, 215Compressibility, 29Computational fluid dynamics, 6Constant-power propulsion, 69, 132Constant-thrust propulsion, 69, 132Control,6

of altitude, 277closed loop, 259derivatives, 228equations, 206force, 9,12

gadgets, 64gradient, 51per g, 60to trim, 48

lareral,205,207,280longitudinal, 33, 204, 207open loop, 204reversal, 64roll, 86, 291vector, 104,228of yaw, 80

Control-free maneuver point, 63Controls:

displacement, 88lateral, 207longitudinal, 207power, 11power boost, 49primary, 204rate, 88

Convergence, 162Conversion factors, 367Convolution, 142Convolution integral, 309Cooper and Harper, IICross-flow, 25Cross product, 95Cycles to double (or halt), 163

Damping:ratio, 164, 266in roll, 150in yaw, 154

Davies, 70Degree of freedom, 6Delta wing, 23

377

Page 386: Stability Control EtkinReid

378 Index

Derivatives:aeroelastic, 156CLa, 141,319CLq,62CI13,83, 148, 340Cia' 321C11le,34Clp' 150, 345Clr' 154,349Cma' 29,141Cmlle,34Clllq'62, 135, 340Cmu' 131, 134Cn13'148,340Cop' 151, 345Cnr' 154, 349CTu,132CXu'131, 133CYp,149CYr' 154Cy13,148CZq,135CZu,131, 133control, 244

jet airplane, 243nondimensional, 207

cross, 151jet transport, 166, 188Lp,120Lr,120i., 120lateral, 156longitudinal, 155nondimensional, general aviation airplane,

238quasistatic, 138Zq,1192.,118Zv,,119'4,119

det,36Determinant, 304Dihedral:

lateral,83longitudinal, 27

Dirigible, 1Disposable load, 74Divergence, 7, 162Dowell,73Downwash, 26, 65, 332

lag of, 147Drag, 19Dutch Roll, 13

approximation, 251

Dynamic gain, 217Dynamic pressure, 23, 30Dynamics of flight, 18

Earth curvature, 3EAS, 37Eigenvalue, 161,210

computation of, 315of jet transport, 166

Eigenvector, 161computation of, 315jet transport, 167-169

Elastic degrees of freedom, 120Elevator:

angle per g, 60, 242angle to trim, 35, 38gearing, 49

Ellis, 70Equations of motion:

Euler's, 100general, 104lateral, 113linear, 111longitudinal, 112nondimensional, 117state vector form, 114

Equilibrium, 6, 20states, 18

Equivalent airspeed, 37Etkin,3,141Euler angle rates, 100Euler angles, 98Euler's equations of motion, 100Extreme value theorems, 308

Feedback,260Feedforward, 298Feel, synthetic, 48Filotas, 147Fin, 86Fixed control, 44Flaps, 23, 64Flax, 20Flexible fuselage, 73, 134Flight, 1Flight path, 99

jet transport, 168Flight simulator, 6Flutter, 72Flying wing, 22Forced oscillation, 146Fourier transform, inverse of, 211Frame of reference, 15Free elevator, 44

Page 387: Stability Control EtkinReid

Index 379

Free-flight model, 6Free oscillation, 146Free rudder, 81Frequency:

circular, undamped, 164reduced, 138, 145response, 214, 243

Gain margin, 265Gates, 3General aviation airplane, 237Generalized inertia, 123Generalized force, 158Giesing, 147Glauert,3Gliding flight, 132Gravity, 3, 367Ground effect, 74Groundspeed,15Gust alleviation, 295Gyrostabilized, 256

normal force, 70pitch damping of, 141

Jet transport, 1lateral modes, 186longitudinal modes, 165

Jones, 3

Kleinman, 11Krendel,11Kuethe and Chow, 133Kuhn, 70

Handling qualities, 5, 11Heat transfer, 3Heaviside theorem, 307Henson, 23Heppe and Celinker, 256High-lift devices, 64High-wing airplane, 85Hinge moment, 41,324

rudder, 81Hughes, 3Human pilot, 5, 7, 8Hunsaker, 3Hypersonic, 20

Lagrange, 3Lagrange's equations, 122Lanchester, 3, 172Landing, 74Laplace, 3Laplace transform, 208, 304, 305

inverse of, 306Lateral aerodynamics, 76Lateral steady states, 237Lawrence, 20Leading-edge suction, 152Lift, 19,23Lift-curve slope, 25, 31Lilienthal, 33Linear air reactions, 109Linear algebra, 303Linearization, 108Load factor, 230, 240Locus, transfer function, 223Logarithmic decrement, 164Longitudinal forces, 19Low-wing airplane, 85

Impulse, 207Impulse response, 210Induced velocity, 25Inertia force, 124Inertial coupling, 255Inertia matrix, 97Instability:

dynamic, 7static, 7

Integral control, 267Interference, 26Interference effects, 25Inverse matrix, 209Inverse problems, 4, 107Inversion theorem, 210

Maneuverability, 60Maneuvers, 8Mass ratio, 61Mathematical model, 93Matrix, 303

adjoint, 304inverse, 304, 308system, 161

McCormick, 19McRuer,llMean aerodynamic center, 356Mean aerodynamic chord, 356Miele, 19Miles, 146Millikan, 70Modes:

effect of CG location, 182effect of density gradient, 180effect of speed and altitude, 1717

Jet:engine, 70induced inflow, 72

Page 388: Stability Control EtkinReid

380 Index

Modes: (continued)divergent, 162jet transport, Dutch Roll, 190jet transport, effect of flight condition, 176jet transport, roll convergence, 189jet transport, spiral, 188lateral, approximations, 193lateral, effect of speed and altitude on, 191longitudinal, approximations, 171longitudinal, of jet transport, 165natural, 162oscillatory, 162phugoid, 166phugoid approximation, 172short-period, 166short-period approximation, 174

Moment of momentum, 95Motion, 1

lateral, 18longitudinal, 18

Moving frame of reference, 316

Nacelles, 25Neutral point, 12, 29

effect of bodies on, 334elevator free, 45flight determination, 40

Newton, 3Newton's laws, 93Nichols diagram, 265Nondimensional system, 115, 116Nonlinear effects, 206Normal modes, 121Nyquist diagram, 217

Oscillating wings, 144

Parabolic polar, 68Partial fractions, 306Performance, 5Period, 163Phase margin, 265Phillips, 256Phugoid:

approximation, 231oscillation, 166, 205

Pinsker, 256Pitch attitude controller, 266Pitching moment, 23

of the body and nacelles, 25of propulsive system, 28of the tail, 26of the wing, 24

Pitch stiffness, 18,21,29

Plant identification, 4Poles, 210, 217Prandtl-Glauert rule, 133Priestly, 70Principal axes, 102Propeller, 67,69

effects, 336fin effect, 80normal force, 80slipstream, 70

Proportional control, 267Propulsive system, 28

effect on pitch stiffness, 66effect on trim, 66

Quasistatic deflections, 121Quasisteady flow, 156

Rate control, 267Reference flight condition, 132Reference steady state, 108Reid,13Resonance, 226Response:

to control, 5to elevator, 229initial, 204longitudinal, 228quasistatic, 227steady state, 204of systems, 207to the throttle, 235transient, 204, 252to turbulence, 5, 295

Reversal of slope, 39Ribner,79Rigid-body equations, 93Rodden, 147Roll damper, 281Roll stiffness, 81Root locus, 267, 281Rotational stiffness, 76Rotation matrix, 99Rotors, 103Routh's criteria, 164Rudder forces, 9, 81Runge-Kutta integration, 206

SAS,7Scalar product, 303Schlichting, 19Sensors, 263Servomechanism, 260Shock tube, 6

Page 389: Stability Control EtkinReid

Index 381

Short-period approximation, 234Sideslip, 237 .Sideslip angle, 17, 77Sidewash angle, 78Sleeman, 70Slipstream effects, 336Slots, 64Small-disturbance theory, 107Smelt, 70Speed controller, 270Speed stability, 177,255Spinning, 12Spiral mode, 12Spiral/roll approximation, 248Spoilers, 87Stability, 6

axes, 102augmentation system(SAS), 7, 76boundary, 75closed loop, 264control-fixed,7control-free, 7derivatives, 129dynamic, 12inherent, 7margin, 40of small disturbances, 5static, 12

limit, 40synthetic, 7of uncontrolled motion, 161weathercock, 77

Stability derivatives:dimensional, 118nondimensional, 117

Stalling, 12State vector, 114, 161Static gain, 214Static longitudinal stability:

general theory, 175static margin, 29static stability criterion, 176static stability limit, 40

Step response, 212, 231Stick movement, 10STOL airplane, 70

basic data, 185longitudinal characteristics, 184

Stringfellow, 23Structure flexibility, 72Submarine, 1Swept-back wing, 23System:

first-order, 219

high-order, 209linear invariant, 206, 207matrix, 161second-order, 224theory, 4, 5

Tab effectiveness, 330Tabs, 47

geared, 48servo, 48spring, 48trim, 47

Tail, 26efficiency factor, 27volume, 28

Tailless aircraft, 32, 35, 36, 46, 63Take-off,74Theodorsen function, 145Thrust, 29

coefficient, 30, 129line, 129vector, 19

Time constant, 211Time to double (or half), 163Tobak, 142Trajectories, 5Transfer functions, 208

closed loop, 281, 271open loop, 264

Transformation:of coordinates, 310of a derivative of a vector, 313of inertias, 351matrix, 311of a matrix, 315of stability derivatives, 354of a vector, 310

Transient states, 8Trim curves, 30Trim-slope criterion, 30Trimmed lift curve slope, 36Truckenbrodt,19Truitt, 20Turbulence, atmospheric, 8Tum, steady, 238

Unit step, 207Unit vectors, 303Upwash,71USAF, 20

Vector product, 303Vectors, 303Vehicle, 1

Page 390: Stability Control EtkinReid

382 Index

Vertical tail, 78volume, 79

Vibration mode, 156Virtual displacement, 124Vortex system, 25Vorticity, 65

Wind tunnel, 6Wing:

bending, 157divergence, 72sweep, 86wake, 26

Wright brothers, 3,23Weight coefficient, 61Wen, 70White noise, 207Wind,16

downburst, 197effects, 196gradient, 197, 199turbulence, 196

Yaw damper, 281, 287Yawing moment, 78Yaw stiffness, 77

Zero-lift line, 31Zeros, 217


Recommended